content
stringlengths 1
15.9M
|
---|
\section{Introduction}
Born and Infeld introduced a nonlinear theory of the electromagnetic field \cite{borninf} with the aim to solve the problem of the infinite self energy of a charged point particle in Maxwell electrodynamics. The spherically symmetric solution for Einstein gravity coupled to Born--Infeld electrodynamics was first obtained by Hoffmann \cite{hoffmann}. This solution does not describe the electron, but it corresponds to a black hole. The field equations of Einstein gravity coupled to Born--Infeld electrodynamics have the vacuum spherically symmetric solution \cite{gibbons,breton}
\begin{equation}
ds^{2}=-f(r)dt^{2}+f^{-1}(r)dr^{2}+r^{2}(d\theta^2+\sin^2\theta d\varphi^2),
\label{bi1}
\end{equation}
with
\begin{equation}
f(r)=1-\frac{2M}{r}+\frac{2}{3b^{2}}\left\{ r^{2}-\sqrt{r^{4}+b^{2}Q^{2}}+
\frac{\sqrt{|bQ|^{3}}}{r}F\left[ \arccos\left( \frac{r^{2}-|bQ|}{r^{2}+|bQ|}
\right) ,\frac{\sqrt{2}}{2}\right]\right\} ,
\label{bi2}
\end{equation}
where $F(\gamma ,k)$ is the elliptic integral of the first kind. As usual, $M>0$ is the ADM mass and $Q^{2}=Q_{E}^{2}+Q_{M}^{2}$ is the square of the charge. The corresponding electric and magnetic inductions are $D(r)=Q_{E}/r^{2}$ and $B(r)=Q_{M}\sin \theta$. With the units adopted ($c=G=1$), $M$, $Q$ and $b$ have dimensions of length. The parameter $b$ measures the departure from Maxwell theory. In the limit $b\rightarrow 0$, the Reissner--Nordstr\"{o}m metric $f(r)=1-2M/r+Q^2/r^2$ is obtained. The geometry given by Eqs. (\ref{bi1}) and (\ref{bi2}) is also asymptotically Reissner--Nordstr\"{o}m for large values of $r$ and it is singular at the origin \cite{breton}. The zeros of $f(r)$ correspond to the horizons. For a given value of $b$, when $0\le |Q|\le Q_d$, the function $f(r)$ has only one zero corresponding to a regular event horizon. For $Q_d<|Q|<Q_c$, $f(r)$ has two zeros; then, an inner horizon and an outer regular event horizon exist. When $|Q|=Q_c$, there is one degenerate horizon. Finally, if $|Q|>Q_c$, the function $f(r)$ has no zeros and there is a naked singularity. The values of $|Q|$ where the number of horizons change, which result from the condition $f(r_{h})=f'(r_{h})=0$, are increasing functions of $|b|/M$.
The Darmois--Israel formalism \cite{daris}, in which the Lanczos equations relate the geometry at both sides of a surface with the induced energy-momentum tensor on it, has become the main tool for the study of the dynamics of highly symmetric layers. The stability of spherical shells was studied by several authors (see \cite{poi,sh1,sh2,sh3} and the references included there). The formalism was applied to bubbles, shells around stars and black holes, and to thin-shell wormholes (see for example \cite{wh1,wh2,wh4} and references therein). Here we present a recent study \cite{eisi11} of the mechanical stability of spherical shells under perturbations preserving the symmetry within the framework of Einstein--Born--Infeld theory. We first introduce the formalism and perform the mathematical construction of shells. Then, we analyze charged bubbles and charged layers around non charged black holes in the framework of Einstein gravity coupled to Born--Infeld electrodynamics.
\section{Charged shells: construction and stability}
We start from two manifolds $\mathcal{M}_1$ and $\mathcal{M}_2$ with metrics of the form (\ref{bi1}) in coordinates $X_{1,2}^\mu=(t_{1,2},r_{1,2}, \theta, \varphi)$, with in general different arbitrary functions $f_{1,2}$, respectively. We cut and paste them at the spherical surface $\Sigma$ defined by $r_{1,2}=a$. The resulting manifold ${\cal M}$ is given by the union of the inner part ($r_1\leq a$) of $\mathcal{M}_1$ and by the outer part ($r_2\geq a$) of $\mathcal{M}_2$ . The line element is continuous across $\Sigma$ if the coordinates at each side are set to satisfy $f_1(a)dt_1^2=f_2(a)dt_2^2$, as required by the thin-shell formalism \cite{daris}. We let the radius $a$ to be a function of the proper time $\tau$ measured on the surface. The coordinates at the surface $\Sigma$ are $\xi ^i=(\tau , \theta,\varphi )$. The relation between the geometry and the shell matter is given \cite{daris} by the Lanczos equations
\begin{equation}
-[K_{ij}]+g_{ij}[K]=8\pi S_{ij},
\label{leq}
\end{equation}
where $g_{ij}$ is the induced metric on $\Sigma$, $K_{ij}$ is the extrinsic curvature, $K$ is its trace, and $S_{ij}$ is the surface energy-momentum tensor; the brackets denote the jump of a given quantity $q$ across the surface: $[q]=q^2\vert_\Sigma -q^1\vert_\Sigma$. If $[K_{ij}]\neq 0$ we have a thin shell at $\Sigma$. The Lanczos equations give the energy density and the pressure at the shell
\begin{equation}
\sigma=-S_\tau^\tau= -\frac{1}{4\pi a}\left(\sqrt{f_2(a)+{\dot a}^2}-\sqrt{f_1(a)+{\dot a}^2}\right),\label{00}
\end{equation}
\begin{equation}
p=S_\theta^\theta=S_\varphi^\varphi=-\frac{\sigma}{2}+\frac{1}{16\pi}\left(\frac{2\ddot a+f'_2(a)}{\sqrt{f_2(a)+{\dot a}^2}}-\frac{2\ddot a+f'_1(a)}{\sqrt{f_1(a)+{\dot a}^2}}\right),
\label{pres}
\end{equation}
where a prime represents $d/dr$, and a dot stands for $d/d\tau $. The equations above, or any of them plus the conservation equation
$d(a^2\sigma)/d\tau+p da^2/d\tau=0$, determine the evolution of the shell. We consider small perturbations preserving the symmetry around a static solution. Our procedure is similar to the treatment in Refs. \cite{poi,sh1,sh2,sh3,wh1}. Provided the equation of state $p=p(\sigma)$, the conservation equation can be formally integrated \cite{wh1} to give $\sigma=\sigma(a)$. From Eq. (\ref{00}) we obtain
\begin{equation}
{\dot a}^2+V(a)=0,
\label{energy}
\end{equation}
where
\begin{equation}
V(a)=S-\frac{1}{4}\left( \frac {m}{a}\right)^2-\left( \frac{a}{m}\right)^2 R^2 ,\label{potential}
\end{equation}
with $S(a)=\left( f_1(a)+f_2(a)\right) /2$, $R(a)=\left( f_1(a)-f_2(a)\right) /2$ and $m(a)=4\pi a^2\sigma$. An equilibrium radius $a_0$ satisfies $V(a_0)=0$ and $V'(a_0)=0$, and stability requires $V''(a_0)>0$. After evaluating the derivatives, the condition for stability gives
\begin{equation}
\frac{m}{4a_0}\left(\frac{m}{a_0}\right)''+ \frac{a_0}{m}\left(\frac{a_0}{m}\right)''R^2 < \Omega(a_0)-\Gamma^2(a_0),
\end{equation}
where
\begin{equation}
\Gamma(a_0) = \frac{a_0}{m}\left[ S'-2\frac{a_0}{m}\left(\frac{a_0}{m}\right)'R^2-2\left(\frac{a_0}{m}\right)^2 RR'\right]
\end{equation}
and
\begin{equation}
\Omega(a_0) = \frac{S''}{2}-\left[\left(\frac{a_0}{m}\right)'\right]^2R^2 -4\frac{a_0}{m}\left(\frac{a_0}{m}\right)'RR' - \left(\frac{a_0}{m}\right)^2\left[ R'^2+RR''\right].
\end{equation}
Here $m$, $S$ and $R$ are given as functions of $a_0$. Using the conservation equation and defining $\eta = p'(a_0)/\sigma'(a_0)$ the condition for stable equilibrium can be put in the form
\begin{equation}
2\left( \frac{a_0}{m} \right) ^4 \left[ \left(\frac{m}{a_0}\right)' \right] ^2 R^2 + \frac{1}{a_0}\left[ \frac{m}{4a_0} - \left( \frac{a_0}{m}\right) ^3 R^2 \right] \left[ \frac{m}{a_0^2}-\left(\frac{m}{a_0}\right)'\right] \left( 1+2\eta \right)<\Omega(a_0)-\Gamma^2(a_0).
\label{sta}
\end{equation}
The subsequent study is carried out in terms of the parameter $\eta $, which for $0<\eta\leq 1$ can be understood as the square of the velocity of sound on the shell. The results above are also valid for wormholes if the outer parts of $\mathcal{M}_1$ and $\mathcal{M}_2$ are taken, and the $(-)$ sign inside the parenthesis in Eqs. (\ref{00}) and (\ref{pres}) is changed by a $(+)$ sign. In particular, if $\mathcal{M}_{1,2}$ are equal copies of the geometry (\ref{bi1}) with $r\geq a$ the results of Refs. \cite{wh1,wh2,wh4,risi09} can be recovered.
\begin{figure}[t!]
\centering
\begin{minipage}{.47\textwidth}
\centering
\includegraphics[width=\textwidth]{fig1.eps}
\caption{Stability regions (grey) for charged shells of normal matter around vacuum ($M_1=0$, $Q_1=0$).}
\label{fig1}
\end{minipage}
\hfill
\begin{minipage}{.47\textwidth}
\centering
\includegraphics[width=\textwidth]{fig2.eps}
\caption{Stability regions (grey) for charged shells of normal matter around a black hole ($M_1/M_2=0.5$, $Q_1=0$).}
\label{fig2}
\end{minipage}
\end{figure}
The geometry consists in the inner part of manifold $\mathcal{M}_1$ and the outer part of manifold $\mathcal{M}_2$ joined by the shell $\Sigma $. Our analysis is restricted to normal matter, so that the weak energy condition $\sigma\geq 0$ and $\sigma+p\geq 0$ must be fulfilled. We first consider bubbles, so we take the outer manifold with mass $M_2$ and charge $Q_2$, and the inner one with both vanishing mass and charge. The radius $a_0$ is chosen larger than the horizon radius of the outer manifold (so that the singularity and the event horizon of the original manifold are both removed). In second place, we analyze charged shells around non charged black holes, so $M_1\neq 0$, $Q_1=0$, $M_2\neq 0$ and $Q_2\neq 0$. Besides demanding that $a_0$ is greater than the horizon radius of the outer manifold, we also have to take $a_0>2M_1$. A necessary condition (but not sufficient for charged shells) for fulfilling the weak energy condition is that $0\leq M_1<M_2$. We present the results graphically in Figs. \ref{fig1} and \ref{fig2}; we take $b=1$, which means a large deviation from Maxwell electrodynamics. We do not restrict our analysis to the range $0<\eta \le 1$, in which $\eta^{1/2}$ can be interpreted as the velocity of sound on the shell, though the results within this range are of more physical interest. The figures illustrate the dependence of the stability regions in terms of the parameters and the constant $b$; in particular, for shells surrounding black holes we show the results corresponding to $M_1/M_2=0.5$. For comparison, the non charged shells results of Ref. \cite{poi} are also displayed. The qualitative behaviour is different if the charge is under or beyond the critical value $Q_c$ (from which the horizon of the outer original manifold vanishes). If $Q<Q_c$, stable configurations are possible only for positive $\eta $, while if $Q\geq Q_c$, negative values of $\eta $ are compatible with stability. In all cases, there are values of $a_0/M_2$ for which stable configurations are possible with $0<\eta \leq 1$. Within this range of $\eta $, if the charge is below $Q_c$ the largest interval of $a_0/M_2$ for which the shell is stable corresponds to $\eta =1$, as it was obtained for non charged shells in Ref. \cite{poi}. When the charge is under $Q_c$, shells around black holes present slightly smaller regions of stability than bubbles; and the stability regions become larger as the charge increases (the same happens for spherical thin-shell wormholes with charge \cite{wh2}). If the charge is equal or beyond $Q_c$, for a given $\eta$ bubbles can be stable for smaller radii $a_0/M_2$ than shells around black holes, and for fixed $a_0/M_2$ bubbles are stable for a smaller range of the parameter $\eta $ than shells around black holes.
\section{Summary}
We have studied the stability under perturbations preserving the symmetry for bubbles and shells around non charged black holes within the framework of Einstein--Born--Infeld theory. The presence of the charge seems to enlarge the stability regions for both bubbles and shells around black holes. The stability regions for bubbles appear to be larger than those of shells around black holes if the charge is under the critical value, while the reverse is true for charges above the critical value. Charged layers with $0<\eta\leq 1$ can be stable for suitable values of the parameters. From a different point of view, we note that with small changes in the formalism the stability of thin-shell wormholes in the same theoretical framework has also been studied (see Ref. \cite{risi09}).
\section*{Acknowledgments}
This work has been supported by Universidad de Buenos Aires and CONICET.
|
\section{Introduction}\label{intro}
In many real-life collective decision making situations, the set of
candidates (or alternatives) may vary while the voting process goes
on, and may change at any time before the decision is final: some
new candidates may join, whereas some others may withdraw. This, of
course, does not apply to situations where the vote takes place in a
very short period of time (such as, typically, political elections
in most countries), and neither does the addition of new candidates during the
process apply to situations where the law forbids
new candidates to be introduced after the voting process has started (which,
again, is the case for most political elections). However, there are
quite many practical settings where this may happen,
especially situations where votes are sent by email during an extended
period of time. This is typically the case when making a decision
about the date and time of a meeting. In the course of the process,
we may learn that the room is taken at a given time slot, making
this time slot no longer a candidate. The opposite case also occurs
frequently; we thought the room was taken on a given date and then
we learn that it has become available, making this time slot a new
candidate.
The paper focuses on candidate addition only. More precisely, the
class of situations we consider is the following. A set of voters
have expressed their votes about a set of (initial) candidates. Then
some new candidates declare their intention to participate in the election. The winner will ultimately be
determined using some given voting rule and the voters' preferences over the set of all candidates.
In this class of
situations, an important question arises: {\em who among the
initial candidates can still be a winner once the voters'
preferences about all candidates are known?} This is important in
particular if there is some interest to detect as soon as possible
the candidates who are not possible winners: for instance,
candidates for a job may have the opportunity to apply for different
positions, and time slots may be released for other potential
meetings.
This question is strongly related to several streams of work in
the recent literature on computational social choice, especially the
problem of determining whether the vote elicitation process can be
terminated \cite{ConitzerSandholm02b,Walsh08}; the possible winner
problem, and more generally the problem of applying a voting rule to
incomplete preferences
\cite{KonczakLang05,PiniRVW07,XiaConitzer11,BetzlerDorn09,BetzlerHemmannNiedermeier09}
or uncertain preferences with probabilistic information
\cite{HAKW09}; swap bribery, encompassing the possible winner
problem as a particular case \cite{ElkindFaliszewskiSlinko09};
voting with an unknown set of available candidates \cite{LuBoutilier10};
the control of a voting rule by the chair via adding
candidates; and resistance to cloning---we shall come back to
the latter two problems in more detail in the related work
section.
Clearly, considering situations where new voters are added is a
specific case of voting under incomplete preferences, where
incompleteness is of a very specific type: the set of candidates is
partitioned in two groups (the initial and the new candidates), and
the incomplete preferences consist of complete rankings on the
initial candidates. This class of situations is, in a sense, dual to a class of situations that has been considered more often, namely,
when the set of {\em voters} is partitioned in two groups: those
voters who have already voted, and those who have not expressed their
votes yet. The latter class of situations, while being a subclass of
voting under incomplete preferences, has been more specifically
studied as a {\em coalitional manipulation problem}
\cite{ConitzerSandholmLang07,FaliszewskiProcaccia10}, where the problem is to
determine whether it is possible for the voters who have not voted
yet to make a given candidate win. Varying sets of voters have also
been studied in the context of compiling the votes of a
subelectorate \cite{CLMR09,XiaConitzerAAAI10}: there, one is
interested in summarizing a set of initial votes, while still being
able to compute the outcome once the remaining voters have expressed
their votes.
The layout of the paper is as follows. In Section \ref{background}
we recall the necessary background on voting and we introduce some
notation. In Section \ref{possible} we state the problem formally,
by defining voting situations where candidates may be added after
the votes over a subset of initial candidates have already been
elicited. In the following sections we focus on specific voting
rules and we study the problem from a computational point of view.
In Section \ref{kapp}, we focus on the family of $K$-approval rules,
including plurality and veto as specific subcases, and give a full
dichotomy result for the complexity of the possible winner problem with
respect to the addition of $k$ new candidates; namely, we show that
the problem is {\sf NP}-complete as soon as $K \geq 3$ and $k \geq
3$, and polynomial if $K \leq 2$ or $k \leq 2$. In Section
\ref{borda} we focus on the Borda rule and show that the problem is
polynomial-time solvable regardless of the number of new candidates. We also exhibit a more general family of voting rules, including Borda, for which this result can be generalized. In Section
\ref{hardone} we show that the problem can be hard for some
positional scoring rules even if only one new candidate is added. In
Section \ref{related} we discuss the relationship to the general
possible winner problem, to the control of an election by the chair
via adding candidates, and to candidate cloning. Section
\ref{conclu} summarizes the results and mentions further research
directions.
\section{Background and notation}\label{background}
Let $C$ be a finite set of {\em candidates}, and $N$
a finite set of {\em voters}. The number of voters is denoted by $n$, and the (total) number of candidates
by $m$.
A {\em $C$-vote} (called simply a vote when this is not ambiguous) is a linear order over $C$, denoted by $\succ$ or by $V$.
We sometimes denote votes in the following way: $a \succ b \succ c$ is denoted by $abc$, etc.
An $n$-voter
{\em $C$-profile}
is a collection $P = \langle V_1,\ldots, V_n\rangle$ of $C$-votes.
Let ${\cal P}_C$ be the set of all $C$-votes and therefore ${\cal P}_C^n$ be the set of all $n$-voter $C$-profiles.
We denote by ${\cal P}_C^*$ the set of all $n$-voter $C$-profiles for $n \geq 1$, {\em i.e.}, ${\cal P}_C^* = \cup_{n \geq 1} {\cal P}_C^n$.
A voting rule on $C$ is a function $r$ from ${\cal P}_C^*$ to $C$.
A voting correspondence is a function from ${\cal P}_C^*$ to $2^C\setminus\{\emptyset\}$.
The most natural way of obtaining a voting rule from a voting correspondence is to break
ties according to a fixed priority order on candidates. In this paper, we
do not fix a priority order on candidates (one reason being that the
complete set of candidates is not known to start with), which means
that we consider voting correspondences rather than rules,
and ask whether $x$ is a {\em possible cowinner} for a given profile
$P$. This is equivalent to asking whether there exists a priority
order for which $x$ is a possible winner, or else whether $x$ is a
possible winner for the most favorable priority order (with $x$
having priority over all other candidates).
This is justified in our context by the fact that specifying such a priority order is problematic when we don't know in advance the identities of the potential new candidates.
With a slight abuse of notation we denote voting correspondences by $r$ just as voting rules. Let $r(P)$ be the set of {\em cowinners} for profile $P$.
For $P \in {\cal P}_C^*$ and $x,x' \in C$, let $n(P,i,x)$ be the number of votes in $P$ ranking $x$ in position $i$,
$ntop(P,x) = n(P,1,x)$ the number of votes in $P$ ranking $x$ first, and $N_P(x,x')$ the number of votes in $P$ ranking $x$ above $x'$.
Let $\vec{s} = \langle s_1, \ldots, s_m\rangle$ be a vector of integers such that $s_1 \geq \ldots \geq s_m$ and $s_1 > s_m$.
The scoring rule $r_{\vec{s}}(P)$ induced by $\vec{s}$ elects the candidate(s) maximizing
$S_{\vec{s}}(x,P) = \sum_{i=1}^m s_i \cdot n(P,i,x)$.
If $K$ is a fixed integer then {\em $K$-approval}, $r_K$, is the scoring rule corresponding to the vector $\vec{s_K} = \langle 1,\ldots, 1, 0, \ldots, 0\rangle$ -- with $K$ 1's and $m-K$ 0's. The $K$-approval score $S_{\vec{s_K}}(x,P)$ of a candidate $x$ is denoted more simply by $S_K(x,P)$: in other words, $S_K(x,P)$ is the number of voters in $P$ who rank $x$ in the first $K$ positions, {\em i.e.}, $S_K(x,P) = \sum_{i = 1, \ldots, K} n(P,i,x)$. When $K = 1$, we get the {\em plurality} rule $r_P$, and when $K = m-1$ we get the {\em veto} (or {\em antiplurality}) rule.
The {\em Borda} rule $r_B$ is the scoring rule corresponding to the vector $\langle m-1,m-2,\ldots,0\rangle$.
We now define formally situations where new candidates are added.
\begin{defin}\label{situation}
A {\em voting situation with a varying set of candidates} is a 4-tuple $\Sigma = \langle N, X, P_X, k \rangle$ where $N$ is a set of voters (with $|N| = n$), $X$ a set of candidates, $P_X = \langle V_1,\ldots, V_n \rangle$ an $n$-voter $X$-profile, and $k$ is a positive integer, encoded in unary.
\end{defin}
$X$ denotes the set of initial candidates, $P_X$ the initial profile, and $k$ the number of new candidates.
Nothing is known a priori about the voters' preferences over the new candidates, henceforth their identity is irrelevant and only their number counts. The assumption that $k$ is encoded in unary ensures that the number of new candidates is polynomial in the size of the input. Most of our results would still hold if the number of new candidates is exponentially large in the size of the input, but for the sake of simplicity, and also because, in practice, $k$ will be small anyway, we prefer to exclude this possibility.
Because the number of candidates is not the same before and after the new candidates come in, we have to consider families of voting rules (for a varying number of candidates) rather than voting rules for a fixed number of candidates.
While it is true that for many usual voting rules there is an obvious way of defining them for a varying number of candidates, this is not the case for all of them, especially scoring rules. Still, some natural scoring rules, including plurality, veto, more generally $K$-approval, as well as Borda, are naturally defined for any number of candidates. We shall therefore consider {\em families of voting rules}, parameterized by the number of candidates ($r^m$). We slightly abuse notation and denote these families of voting rules by $r$, and consequently
often write $r(P)$ instead of $r^m(P)$. The complexity results we give in this paper make use of such families of voting rules, where the number of candidates is variable.
If $P$ is a $C$-profile and $C' \subseteq C$, then the projection of $P$ on $C'$, denoted by $P^{\downarrow C'}$, is obtained
by deleting all candidates in $C \setminus C'$ in each of the votes of $P$, and leaving unchanged the ranking on the candidates of $C'$.
For instance, if $P = \langle abcd, dcab \rangle$, then $P^{\downarrow \{a,b\}} = \langle ab, ab\rangle$ and
$P^{\downarrow \{a,b,c\} }= \langle abc, cab\rangle$.
In all situations, the set of initial candidates is denoted by $X =
\{x_1,\ldots, x_p\}\cup \{x^*\}$, the set of the $k$ new candidates
is denoted by $Y = \{y_1, \ldots, y_k\}$. If $P_X$ is an $X$-profile
and $P'$ an $X \cup Y$-profile, then we say that $P'$ extends $P_X$
if the projection of $P'$ on $X$ is exactly $P_X$.
For instance, let $X = \{x_1,x_2,x_3\}\cup \{x^*\}$, $Y =
\{y_1,y_2\}$; the profile $P'= \langle x_1 y_1 x^* x_2 y_2 x_3, y_1
y_2 x_1 x_2 x_3 x^*, x_3 x_2 y_2 x^* y_1 x_1\rangle$ extends the
$X$-profile $P_X= \langle x_1 x^* x_2 x_3, x_1 x_2 x_3 x^*, x_3 x_2
x^* x_1\rangle$.
\section{Possible winners when new candidates are added}\label{possible}
We recall from \cite{KonczakLang05} that given a collection $\langle
P_1, \ldots, P_n\rangle$ of partial strict orders on $C$
representing some incomplete information about the votes, a
candidate $x^*$ is a possible winner if there is a profile $\langle
T_1, \ldots, T_n\rangle$ where each $T_i$ is a ranking on $C$
extending $P_i$ in which $x^*$ wins. Reformulated for the case where
$P_i$ is a ranking of the initial candidates (those in $X$), we get
the following definition:
\begin{defin}\label{posswin
Given a voting situation $\Sigma = \langle N, X, P_X, k \rangle$,
and a collection $r$ of voting rules, we say that $x^* \in X$ is a {\em possible cowinner with respect to $\Sigma$ and $r$} if there is a $(X \cup Y)$-profile $P'$ extending $P_X$ such that $x^* \in r(P')$, where $Y = \{ y_1, \ldots, y_k \}$ is a set of $k$ new candidates.
\end{defin}
Note that we do not have $Y$ in the input, because it would be redundant with $k$: it is enough to know the {\em number} of new candidates. Note also that all new candidates $\{y_1,\ldots, y_k\}$ have to appear in the extended votes composing $P'$.
Also, we do not consider the problem of deciding whether a new candidate $y_j$ is a possible cowinner, because it is trivial. Indeed, as soon as the voting correspondence
satisfies the extremely weak property that a candidate ranked first by all voters is always a cowinner (which is obviously satisfied by all common voting rules), any new candidate is a possible cowinner.
We now define formally the problems we study in this paper.
\begin{defin}
Given a collection $r$ of voting rules, the {\sc possible cowinner problem with new candidates} (or {\sc PcWNC}) for $r$ is defined as follows:
\begin{description}
\item[Input] A voting situation $\Sigma = \langle N, X, P, k\rangle$ and a candidate $x^*\in X$.
\item[Question] Is $x^*$ a possible cowinner with respect to $\Sigma$ and $r$?
\end{description}
Also, the subproblem of {\sc PcWNC} where the number $k$ of new candidates is fixed will be denoted by {\sc PcWNC}$(k)$.
\end{defin}
We can also define the notion of {\em necessary cowinner} with
respect to $\Sigma$ and $r$: $x^* \in X$ is a {\em necessary
cowinner with respect to $\Sigma$, $Y$, and $r$} if for every $(X
\cup Y)$-profile $P'$ extending $P_X$ we have $x^* \in r(P)$.
However, the study of necessary cowinners in this particular setting
will almost never lead to any significant results.
There may be necessary cowinners among the
initial candidates, but this will happen rarely (and this case
will be discussed for a few specific voting rules in the
corresponding parts of the paper).
Now we are in position to consider specific voting rules.
\section{$K$-approval}\label{kapp}
As a warm-up we start by considering the plurality
rule.
\subsection{Plurality}\label{pluve}
Let us start with an example: suppose $X = \{a,b,c\}$, $n = 13$, and the plurality scores in $P_X$ are $a \mapsto 6$, $b \mapsto 4$, $c \mapsto 3$.
There is only one new candidate ($y$). We have:
\begin{enumerate}
\item
$a$ is a possible cowinner ($a$ will win in particular if the top candidate of every voter remains the same);
\item $b$ is a possible cowinner: to see this, suppose that 2 voters who had ranked $a$ first now rank $y$ first; the new scores are $a \mapsto 4$, $b \mapsto 4$, $c \mapsto 3$, $y \mapsto 2$;
\item $c$ is not a possible cowinner: to reduce the scores of $a$ (resp. $b$) to that of $c$, we need at least 3 (resp. 1) voters who had ranked $a$ (resp. $b$) first to now rank $y$ first; but this then means that $y$ gets at least 4 votes, while $c$ has only 3.
\end{enumerate}
More generally, we have the following result:
\begin{prop}\label{prop-plura}
Let $P_X$ be an $n$-voter profile on $X$, and $x^* \in X$. The candidate $x^*$ is a possible cowinner for $P_X$ and plurality
with respect to the addition of $k$ new candidates
if and only if
$$ntop(P_X,x^*) \geq \frac{1}{k} \cdot \sum_{x_i \in X}\max(0, ntop(P_X,x_i)-ntop(P_X,x^*))$$
\end{prop}
\noindent {\em Proof:~}
Suppose first that the inequality holds.
We build the following $(X \cup Y)$-profile $P'$ extending $P_X$:
\begin{enumerate}
\item for every candidate $x_i$ such that $ntop(P_X,x_i) > ntop(P_X,x^*)$ we simply take $ntop(P_X,x_i) - ntop(P_X,x^*)$
arbitrary votes ranking $x_i$ on top and place one of the $y_j$'s on top of the vote (and the other $y_j$'s anywhere),
subject to the condition that no $y_j$ is placed on top of a vote more than $ntop(P_X,x^*)$ times. (This is possible because the inequality is satisfied).
\item in all other votes (those not considered at step 1), place all $y_j$'s anywhere except on top.
\end{enumerate}
We obtain a profile $P'$ extending $P_X$. First, we have $ntop(P', x^*) = ntop(P_X, x^*)$,
because in all the votes in $P_X$ where $x^*$ is on top, the new top candidate in the corresponding vote in $P'$ is still $x^*$
(cf. step 2), and all the votes in $P_X$ where $x^*$ was not on top obviously cannot have $x^*$ on top in the corresponding vote in $P'$.
Second, let $x_i \neq x^*$. If $ntop(P_X,x_i) \leq ntop(P_X,x^*)$ then $ntop(P',x_i) = ntop(P_X,x_i)$; and
if $ntop(P_X,x_i) > ntop(P_X,x^*)$ then we have $ntop(P',x_i) = ntop(P_X,x_i) - (ntop(P_X,x_i) - ntop(P_X,x^*)) = ntop(P_X,x^*)$.
Therefore, \jl{$x^*$ is a cowinner for plurality in $P'$}.\\
Conversely, if the inequality is not satisfied, in order for $x^*$ to become \jl{a cowinner} in $P'$,
the other $x_i$'s must lose globally an amount of $\sum_{x_i \in X}\max(0, ntop(P_X,x_i)-ntop(P_X,x^*))$ votes.
But since we have $\sum_{x_i \in X}\max(0, ntop(P_X,x_i)-ntop(P_X,x^*)) > k \cdot ntop(P_X,x^*)$, for at least one of the $y_j$'s
it must hold that $ntop(P',y_j) > ntop(P',x^*)$; therefore $x^*$ cannot be \jl{a cowinner} for plurality in $P'$.
\hfill \qed\\
We do not need to pay much attention to the veto rule, since the characterization of possible cowinners is
trivial. Indeed, by placing any of the new candidates below $x^*$ in every vote of $P_X$ where $x^*$ is ranked at the bottom position, we obtain a vote $P'$ where no one vetoes $x^*$, so any candidate is a possible cowinner.
As a corollary, computing possible cowinners for the rules of plurality (and veto) with respect to candidate addition can be computed in polynomial time (which we already knew, since possible cowinners for plurality and veto can be computed in polynomial time \cite{BetzlerDorn09}).
\subsection{$K$-approval, one new candidate}\label{kapp-1}
We start with the
case where a single candidate is added.
Recall that we denote by $S_K(x_j,P_X)$ the score of $x_j$ for $P_X$
and $K$-approval (\emph{i.e.} the number of voters who rank $x_j$
among their top $K$ candidates); and by $n(P_X,K,x_j)$ the number of
voters who rank $x_j$ exactly in position $K$.
\begin{prop}\label{prop-kapp}
Let $K$ be an positive integer,
$P_X$ be an $n$-voter profile on $X$, and $x^* \in X$. The candidate $x^*$ is a possible cowinner for $P_X$
and $K$-approval with respect to the addition of one new candidate
if and only if the following two conditions hold:
\begin{enumerate}
\item for every $x_i \neq x^*$, if $S_K(x_i,P_X) > S_K(x^*,P_X)$ \\
then $n(P_X,K,x_i) \geq S_K(x_i,P_X) - S_K(x^*,P_X)$.
\item {\small
$S_K(x^*,P_X) \geq \sum_{x_i \in X}\max(0, S_K(x_i,P_X) - S_K(x^*,P_X))$}
\end{enumerate}
\end{prop}
\noindent {\em Proof:~}
Assume conditions (1) and (2) are satisfied. Then, we build the following $(X \cup \{y\})$-profile extending $P_X$:
\begin{itemize}
\item[(i)] for every $x_i$ such that $S_K(x_i,P_X) > S_K(x^*,P_X)$, we take $S_K(x_i,P_X) - S_K(x^*,P_X)$
arbitrary votes who rank $x_i$ in position $K$ in $P_X$ and place
$y$ on top (condition (1) ensures that we can find enough such votes).
\item[(ii)] in all other votes (those not considered at step (i)), place $y$ in the bottom position.
\end{itemize}
We obtain a profile $P'$ extending $P_X$. First, we have
$S_K(x^*,P') = S_K(x^*,P_X)$, because (a) all votes in $P_X$ ranking
$x^*$ in position $K$ are extended in such a way that $y$ is placed
in the bottom position, therefore $x^*$ gets a point in each of
these votes if and only if it got a point in $P_X$, and (b) in all the
other votes (those where $x^*$ is not ranked in position $K$ in
$P_X$), $x^*$ certainly gets a point in $P'$ if and only if they
got a point in $P_X$. This holds both in the case where $y$ was added at the
top or the bottom of the vote. Second, for every $x_i$ such that
$S_K(x_i,P_X) > S_K(x^*,P_X)$, $x_i$ loses exactly $S_K(x_i,P_X) -
S_K(x^*,P_X)$ points when $P_X$ is extended into $P'$, therefore
$S_K(x_i,P') = S_K(x_i,P_X) - S_K(x_i,P_X) + S_K(x^*,P_X) =
S_K(x^*,P_X)$. Third, $S_K(y,P') = \sum_{x_i \in X}\max(0,
S_K(x_i,P_X) - S_K(x^*,P_X)) \leq S_K(x^*,P_X)$---because of
(2)---hence $S_K(y,P') \leq S_K(x^*,P')$. Therefore, $x^*$ is a
cowinner for $K$-approval in $P'$.
Now, assume condition (1) is not satisfied, that is, there is an
$x_i$ such that $S_K(x_i,P_X) > S_K(x^*,P_X)$ and such that $n(P_X,
K,x_i) < S_K(x_i,P_X) - S_K(x^*,P_X)$. There is no way of having
$x_i$ lose more than $S_K(x_i,P_X)$ points, therefore $x^*$ will
never catch up with $x_i$'s advantage and is therefore not a possible
\jl{cowinner}. Finally, assume condition (2) is not satisfied, which means
that we have $\sum_{x_i \in X}\max(0, S_K(x_i,P_X) - S_K(x^*,P_X)) >
S_K(x^*,P_X)$. Then, in order for $x^*$ to reach the score of
$x_i$'s we must add $y$ in one of the top $K$ positions in a number
of votes exceeding $S_K(x^*,P_X)$, therefore $S_K(y,P')
> S_K(x^*,P_X) \geq S_K(x^*,P')$, and therefore $x^*$ is not a
possible cowinner.
\hfill \qed\\
Therefore, computing possible cowinners for $K$-approval with respect to the addition of {\em one} candidate can be done in polynomial time.
\subsection{$2$-approval, any (fixed) number of new candidates}\label{2app}
For each profile $P$ and each candidate $x'$, we simply write $s(x',P)$ for the
score of $x'$ in $P$ under $r_2$, that is,
$s(x',P)=S_2(x',P)$, \emph{i.e.} the number of times that $x'$ is
ranked within the top two positions in $P$.
Let $P_X = \langle V_1, \ldots, V_n\rangle$ be an initial profile and $Y = \{y_1,\ldots, y_k\}$ the set of new candidates. Let $x^* \in X$. We want to know whether $x^*$ is a possible cowinner for 2-approval and $P_X$.
Let us partition $P_X$ into $P_1$, $P_2$ and $P_3$, where $P_1$
consists of the votes in which $x^*$ is ranked in the top position, $P_2$ consists of the votes in which $x^*$ is ranked in the second position and $P_3$ consists of the votes in
which $x^*$ is {\em not} ranked within the top two positions.
Let $P$ be an extension of $P_X$ to $X \cup Y$.
For each candidate $x'\in X$, we define the following three subsets of $P$:
\begin{itemize}
\item $\text{HP}(P,x')$ is the set of votes in $P$ where $x'$ is ranked in the second
position and neither $x^*$ nor any new candidate is ranked in the top position (HP stands for
``high priority'').
\item $\text{MP}(P,x')$ is the set of votes in $P$ where $x^*$ or any new candidate is ranked in the top position and $x'$ is ranked in the second position (MP stands for ``medium priority'').
\item $\text{LP}(P,x')$ is the set of votes in $P$ where $x'$ is ranked in the top position and some $x'' \in X\setminus\{x^*\}$ is ranked in the second position (LP stands for ``low priority'').
\end{itemize}
These definitions also apply to $P_X$; our definitions then simplify into: $\text{HP}(P_X,x')$ is the set of votes in $P_X$ where $x'$ is ranked second and $x^*$ is not ranked first; $\text{MP}(P_X,x')$ is the set of votes in $P_X$ where $x^*$ is ranked first and $x'$ is ranked second; $\text{LP}(P_X,x')$ is the set of votes in $P_X$ where $x'$ is ranked first and $x^*$ is not ranked second.
These definitions are summarized in Figure \ref{HPMPLP}.
Finally, for $x\in X\cup Y$, let $\Delta(P,x)=S_2(x,P)-S_2(x^*,P)$.
\begin{figure}[t]
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
& top candidate belongs to & $2^{nd}$ candidate belongs to \tabularnewline
\hline
$\text{HP}(P,x')$ & $X\backslash\{x^{*}\}$ & $\{x'\}$\tabularnewline
\hline
$\text{MP}(P,x')$ & $Y\cup\{x^{*}\}$ & $\text{\{}x'\}$\tabularnewline
\hline
$\text{LP}(P,x')$ & $\{x'\}$ & $X\backslash\{x^{*}\}$\tabularnewline
\hline
\end{tabular}
\end{center}
\caption{A vote $V\in P$ belongs respectively to the sets $\text{HP}(.),\text{MP}(.),\text{LP}(.)$ if its top two candidates belong to the respective sets.
} \label{HPMPLP}
\end{figure}
Let us compute these sets on a concrete example, which will be reused throughout the section.
\begin{example}
Let $X = \{x^*, x_1, \ldots, x_6\}$ and consider the following profile $P_X$ consisting of 19 votes (we only mention the first two candidates in each vote): \\
\begin{footnotesize}
\[
\begin{array}{ccccccccccccccccccc}
v_1 & v_2 & v_3 & v_4 & v_5 & v_6 & v_7 & v_8 & v_9 & v_{10} & v_{11} & v_{12} & v_{13} & v_{14} & v_{15} & v_{16} & v_{17} & v_{18} & v_{19} \\
\hline
x^* & x_1 & x_2 & x_3 & x_1 & x_1 & x_1 & x_2 & x_2 & x_2 & x_2 & x_2 & x_3 & x_3 & x_3 & x_3 & x_3 & x_3 & x_4 \\
x_1 & x^* & x^* & x^* & x_4 & x_4 & x_5 & x_1 & x_3 & x_4 & x_5 & x_5 & x_1 & x_2 & x_4 & x_4 & x_5 & x_6& x_6 \\
\vdots & \vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots
\end{array}
\]
\end{footnotesize}
\Omit{
\begin{tabular}{lllllll}
$v_1: c x_1$ & $v_2: x_1 c$& $v_3: x_2 c$& $v_4: x_3 c$& $v_5: x_1 x_4$& $v_6: x_1 x_4$& $v_7: x_1 x_5$ \\
$v_8: x_2 x_1$& $v_9: x_2 x_3$& $v_{10}: x_2 x_4$& $v_{11}: x_2 x_5$& $v_{12}: x_2 x_5$& $v_{13}: x_3 x_1$\\$v_{14}: x_3 x_2$& $v_{15}: x_3 x_4$& $v_{16}: x_3 x_4$& $v_{17}: x_3 x_5$& $v_{18}: x_3 x_6$& $v_{19}: x_4 x_6$
\end{tabular}\\
}
We have $P_1 = \{v_1\}$, $P_2 = \{v_2, v_3, v_4\}$ and $P_3 = \{v_5, \ldots, v_{19} \}$.
This is summarized together with the priority classification in the following table:
$$\begin{array}{|l|l|l|l|c|}
\hline
& \mbox{{\rm HP}} & \mbox{\rm MP} & \mbox{\rm LP} & \Delta(P_X,x_i) \\ \hline
x_1 & v_8, v_{13} & v_1 & v_5, v_6, v_7 & 3\\
x_2 & v_{14} && v_8, v_9, v_{10}, v_{11}, v_{12} & 3\\
x_3 & v_9 && v_{13}, v_{14}, v_{15}, v_{16}, v_{17}, v_{18} & 4\\
x_4 & v_5, v_6, v_{10}, v_{15}, v_{16} && v_{19} & 2 \\
x_5 & v_7, v_{11}, v_{12}, v_{17} & & & 0\\
x_6 & v_{18}, v_{19} & & & -2\\ \hline
\end{array}$$
\end{example}
If $P^*$ is an extension of $P_X$ to $X \cup Y$ then we write
$P^* = \langle V_1^*, \ldots, V_n^*\rangle$, where $V_i^*$ is the vote over $X \cup Y$
extending $V_i$.
We now establish a useful property of the extensions
of $P_X$ for which $x^*$ is a cowinner.
Without loss of generality,
we assume that in every vote $V_i^*$, every new candidate $y_j$ is ranked either in the first two positions, or below all candidates of $X$.
\begin{prop}
\label{prop:opt} If there exists an extension $P$ of $P_X$ such that
$x^*\in r_2(P)$, then there exists an extension $P^*$ of $P_X$ such that $x^*\in r_2(P^*)$, and satisfying
the following conditions:
\begin{enumerate}
\item For each $V_i \in P_X$, if $x^*$ is ranked within the top two positions in $V_i$, then $x^*$ is also ranked within the top two positions in $V_i^*$.
\item For each $V^*_i\in P^*$, if the top candidate of $V_i^*$ is not in $Y$ then the second-ranked candidate of $V^*_i$ is not in $Y$ either.
\item For each $x'\in X\setminus\{x^*\}$ and each $V_i\in \text{\rm MP}(P_X,x')\cup \text{\rm LP}(P_X,x')$, if $x'$ is not ranked within the top two positions in $V^*_i$,
then for each $V_j\in \text{\rm HP}(P_X,x')$, $x'$ is not ranked within the top two positions in $V_j^*$.
\end{enumerate}
\end{prop}
\beginproo
We consider in turn the different conditions:
\begin{enumerate}
\item This is because if there exists $V'\in P$
such that $x^*$ is not in the
top two positions whereas $x^*$ is in the
top two positions in its original vote $V\in P_X$, then we can simply move all of candidates in
$Y$ ranked higher than $x^*$ to the bottom positions. Let
$V^*$ denote the vote obtained this way. By replacing $V'$ with
$V^*$, we increase the score of $x^*$ by $1$, and the score of each
other candidate by no more than $1$, which means that
$x^*$ is still a cowinner.
\item If there exists $V'\in P$ such that $x'\in X$ is ranked in the
top position and $y\in Y$ is ranked in the second position, then
we simply obtain $V^*$ by switching $y$ and $x'$.
\item The condition states that for each candidate $x'$, whenever we want to reduce
its score, we should first try to reduce it by putting a new
candidate $y\in Y$ on top of some vote in $V\in \text{HP}(P_X,x')$.
This is because by putting $y$ on top of some vote in
$\text{HP}(P_X,x')$, we may use only one extra candidate $y' \in Y$
to reduce by one unit the score of the candidate ranked at the top
position of $V$. Formally, suppose there exist $V_1\in
\text{HP}(P_X,x')$ and $V_2\in \text{MP}(P_X,x')\cup
\text{LP}(P_X,x')$ such that $x'$ is within the top two positions of
$V_1'$ (the extension of $V_1$) but not within the top two positions
of $V_2'$ (the extension of $V_2$). Let $y\in Y$ be any candidate
ranked within the top two positions of $V_2'$. Let $V_2^*$ denote
the vote obtained from $V_2'$ by moving $y$ to the bottom, and let
$V_1^*$ denote the vote obtained from $V_1'$ by moving $y$ to the
top position. Next, we replace $V_1'$ and $V_2'$ by $V_1^*$ and
$V_2^*$, respectively. It follows that the score of each candidate
does not change, which means that $x^*$ is still a cowinner.
We repeat this procedure until statement (3) is
satisfied for every $x'\in X\setminus\{x^*\}$. Since after each
iteration there is at least one additional vote that will never be
modified again, this procedure ends in $O(|P_X|)$ times.
\end{enumerate}
\hfill \qed\\
Proposition~\ref{prop:opt} simply tells us that when looking for an
extension that makes $x^*$ a cowinner, it suffices to restrict our
attention to the extensions that satisfy conditions (1) to (3).
Moreover, using (1) of Proposition~\ref{prop:opt}, we
deduce that $s(x^*,P^*)=s(x^*,P_X)$. Hence, for votes $V\in P_2$
(the votes in which $x^*$ is ranked in the second position), we can
assume that the new candidates of $Y$ are put in bottom positions
in $P^*$.
Define $X^\bullet$ as the set of all candidates in $X$ such that $\Delta(P_X,x_i) > 0$.
Our objective is to reduce all score differences to $0$ for $x\in X^\bullet$, while keeping the score differences
of each new candidate non-positive. (We do not have to care about the candidates in $X \setminus X^\bullet$).
The intuition underlying our algorithm is that
when trying to reduce $\Delta(P,x_i)$ on the current profile $P$, we first try to use the votes in $\text{HP}(P_X,x_i)$, then the votes in $\text{MP}(P_X,x_i)$, and
finally the votes in $\text{LP}(P_X,x_i)$. This is because putting some
candidates from $Y$ in the top positions in the votes of
$\text{HP}(P_X,x_i)$ not only reduces $\Delta(P,x_i)$ by one unit, but also
creates an opportunity to
``pay'' one extra candidate from
$Y$ to reduce $\Delta(P,x_j)$ by one unit, where $x_j$ is the candidate ranked on
top of this vote. For the votes in $\text{MP}(P_X,x_i)$, we can only
reduce $\Delta(P_X,x_i)$ by one unit without any other benefit. For the votes in
$\text{LP}(P_X,x_i)$ we will have to use two candidates from $Y$ to
bring down $\Delta(P,x_i)$ by one unit; however, if we already put some $y\in Y$ in
the top position in order to reduce $\Delta(P,x_j)$, where $x_j$ is the
candidate ranked in the second position in the original
vote, then we only need to pay one extra candidate in $Y$ to
reduce $\Delta(P,x_i)$ by one unit.
Therefore, the major issue consists in finding the most
efficient way to choose the votes in $\text{HP}(P_X,x_i)$ to reduce
$\Delta(P,x_i)$, when $\Delta(P,x_i)\leq |\text{HP}(P,x_i)|$. We will solve this problem by reducing it
to a max-flow problem.
The algorithm is composed of a main function \algoname{CheckCowinner(.)} which comes together with two sub-functions \algoname{AddNewAlternativeOnTop(.)} and \algoname{BuildMaxFlowGraph(.)} that we detail first.
\LinesNumbered
\begin{algorithm}[h]
\caption{AddNewAlternativeOnTop$(P,V,Y)$} \label{algo:addnew}
$y_i\leftarrow argmin_j \left\{ \Delta(P,y_j) : y_j\in Y \right\}$ // take lowest index $i$ when tie-breaking \\
add $y_i$ on top of $V$ and update $P$\\
return $P$\\
\end{algorithm}
The procedure \algoname{AddNewAlternativeOnTop} simply picks new candidates to be put on top of votes, and updates subsequently the profile.
Note that in this procedure, candidates from $Y$ to be added on top of the votes are those with the lowest score (or the lowest index, in case of ties). This results in choosing new candidates in a cyclic order $y_1\rightarrow y_2\ldots\rightarrow y_{|Y|}\rightarrow y_1 \ldots$
As for the function \algoname{BuildMaxFlowGraph}$(P,x^{*},X_1,X_2)$, \nicobf{it} builds the weighted directed graph $G=\langle W,E \rangle$ defined as follows:
\begin{itemize}
\item $W = \{s,t\}\cup X_1\cup X_2\cup \bigcup_{x_i\in X_2} \text{LP}(P,x_i)$; \label{BMFG_startinit}
\item $E$ contains the following weighted edges:
\begin{itemize}
\item for each $x \in X_1$, an edge $(s,x)$ with weight \jm{$\Delta(P,x)$};
\item for each $x \in X_2$ and each $V\in \text{LP}(P,x)$: an edge $(V,x)$ with weight $1$; plus, if the candidate $x'$ in second position in $V$ is in $X_1$, an edge $(x',V)$ \jm{with weight $1$};
\item for each $x \in X_2$, an edge $(x,t)$ with weight $\Delta(P,x)$.
\end{itemize}
\end{itemize}
We refer the reader to Figure \ref{figureflow} for an illustration. (Once this graph is constructed, any standard function to compute a flow $\phi$ of maximal value can of course be used).
We are now in a position to detail the main function \algoname{CheckCowinner(.)}.
\LinesNumbered
\begin{algorithm}[H]
\label{algocc}
\caption{CheckCowinner$(P_X,x^{*},Y)$}
$P\leftarrow P_{X}$ \label{CC_init} \\
$T \leftarrow 0$ // number of calls AddNewAlternativeOnTop \\
$X_{1}\leftarrow\left\{ x_{i} \in X^\bullet:\left|\text{HP}(P_X,x_{i})\right|>\Delta(P_X,x_i)\right\} $ \\
$X_{2}\leftarrow\left\{ x_{i} \in X^\bullet:\left|\text{HP}(P_X,x_{i})\right|\le \Delta(P_X,x_i)\right\} $\label{CC_endinit} \\
$REM \leftarrow \emptyset$ \\
\For{$x_{i}\in X_{2}$ \label{CC_for1}}{
\For{$V\in \text{HP}(P,x_{i})$}{
$P\leftarrow \algoname{AddNewAlternativeOnTop}(P,V,Y)$\\
$T\!+\!+$ \label{CC_endfor1}
}
}
\For{$x_{i}\in X_{2}$ \label{CC_for2}}{
\For{$V\in \text{MP}(P,x_{i})$ \label{CC_for22}}{
\If{$\Delta(P,x_i) > 0$ \label{CC_if1}}{
$P\leftarrow \algoname{AddNewAlternativeOnTop}(P,V,Y)$\\
$T\!+\!+$
}
\Else{$REM \leftarrow REM \cup \{x_i\}$ \label{CC_donewithx2} \label{CC_endfor2}
}
}
}
$X_2 \leftarrow X_2 \backslash REM$ \\
\If{$\exists y\in Y$ \text{such that} $\Delta(P,y)>0$ \label{CC_if2}}
{return {\em false} \label{CC_existsy}}
$G \leftarrow \algoname{BuildMaxFlowGraph}(P,x^*,X_1,X_2)$ \label{CC_build}
\\
$\phi \leftarrow \algoname{ComputeMaxFlow}(G,s,t)$ \label{CC_flow} \\
\If{ $F \ge \sum_{i\leq m-1}\Delta(P,x_i)+\sum_{x_i\in
X_2}\Delta(P,x_i)-(|Y| \cdot s(x^*,P_X)-T)$ \label{CC_if2}}
{return {\em true}}
return {\em false}
\end{algorithm}
\begin{prop}
Given a profile $P_X$ on $X$, a candidate $x^*\in X$ and a set of new candidates $Y$, a call to algorithm \algoname{CheckCowinner}$(P_X,x^*,Y)$
returns in polynomial time the answer \emph{true} if and only if there exists an extension of $P_X$ in which $x^*$ is a cowinner.
\end{prop}
\noindent {\em Proof:~}
Algorithm \ref{algocc} starts by partitioning $X^\bullet$ into $X_1$ and $X_2$: an alternative $x \in X^\bullet$ is in $X_1$ if $|\text{HP}(P_X,x)|>\Delta(P_X,x_i)$ and in $X_2$ if $|\text{HP}(P_X,x)|\le \Delta(P_X,x)$.
Let $x \in X_2$. Then
by item (3) of Proposition~\ref{prop:opt}, for each vote in $V\in \text{HP}(P,x)$, we can safely put one candidate from $Y$ in the top position of $V$; this is done in the first phase of Algorithm \ref{algocc}, lines \ref{CC_for1} to \ref{CC_endfor1}.
Note that after adding a new candidate on top of a vote $V \in \text{HP}(P,x)$ and after updating $P$,
the modified vote will no longer belong to $\text{HP}(P,x)$. Instead, it will now belong to $\text{MP}(P,x')$
for some other candidate $x'$.
When Phase 1 is over, the score of $x \in X_2$ may still need to be lowered down, which can be done next by using votes from $\text{MP}(P_X,x)$. This is what Phase 2 does, from line \ref{CC_for2} to line \ref{CC_endfor2}.
There are three possibilities:
\begin{enumerate}
\item $|\text{HP}(P_X,x)| = \Delta(P_X,x)$. In this case, the votes in $\text{HP}(P_X,x)$ are sufficient to make $x^*$ catch up $x$: after Phase 1, we have $\Delta(P_X,x) = 0$ and Phase 2 is void; we are done with $x$.
\item $|\text{HP}(P_X,x)| < \Delta(P_X,x)$ and $|\text{HP}(P_X,x)| + |\text{MP}(P_X,x)| \geq \Delta(P_X,x)$: in this case, to make $x^*$ catch up $x$, it is enough to take $\Delta(P_X,x)-|\text{HP}(P_X,x)|$ arbitrary votes in $\text{MP}(P_X,x)$ and add one new candidate on top of them; this is what Phase 2 does, and after that we are done with $x$.
\item $|\text{HP}(P_X,x)| + |\text{MP}(P_X,x)| < \Delta(P_X,x)$: in this case, because of Proposition~\ref{prop:opt}, we know that it is safe to add one new candidate on top of {\em all} votes of $\text{MP}(P_X,x)$; this is what Phase 2 does; after that, we still need to lower down the score of $x$, which will require to add new candidates on top of votes of $\text{LP}(P_X,x)$.
\end{enumerate}
If at this point a newly added candidate has a score higher than $x^*$, then $x^*$ cannot win, and we can stop the program (line \ref{CC_existsy}).
For readability, let us denote by $\widetilde{P}$ the profile obtained after Phases 1 and 2. For each
$x \in X_2$ satisfying condition 3, the only way to reduce $\Delta(\widetilde{P},x)$ is to put two
candidates of $Y$ within the top two positions in a vote
of $\text{LP}(\widetilde{P},x)$, because in Phases 1 and 2
we have used up all the votes in $\text{HP}(P,x)$ and $\text{MP}(P,x)$.
Now, reducing $\Delta(\widetilde{P},x)$ by one unit will cost us two candidates in
$Y$, but meanwhile, $\Delta(\widetilde{P},x')$ is also reduced by one unit, where $x'$ is
the candidate ranked in the second position in $V$. We
must have $x'\in X_1$. We note that
$\bigcup_{x\in X_2}\text{LP}(P_X,x)\subseteq\bigcup_{x'\in X_1}\text{HP}(P_X,x')$.
Choosing optimally the votes in $\text{LP}(P_X,x)$ for each
$x\in X_2$ can be done by solving an integral max-flow instance which is build by algorithm \algoname{BuildMaxFlowGraph}
(note that in case where either $X_1$ or $X_2$ is empty, we just assume that the flow has a null value).
Let us show that $x^*$ is a possible cowinner if and only if the value of the flow from $s$ to $t$ is at least $\sum_{i\leq m-1}\Delta(\widetilde{P},x_i)+\sum_{x_i\in X_2}\Delta(\widetilde{P},x_i)-(|Y| \cdot s(x^*,P_X)-T)$.
Observe that the flow does not necessarily bring all $\Delta(\widetilde{P},x_i)$ to 0, therefore we sometimes need a postprocessing
consisting of adding further new candidates on top of some votes
(see steps \nicobf{2 and 3} below).\\
Suppose first that the above max-flow instance has a solution whose value which is at
least $$\sum_{i\leq m-1}\Delta(\widetilde{P},x_i)+\sum_{x_i\in
X_2}\Delta(\widetilde{P},x_i)-(|Y| \cdot s(x^*,P_X)-T)$$
We show how
to solve our cowinner problem from the solution to this flow problem. Because the instance is integral, there must exist an integral solution. We arbitrarily choose one integral solution $\phi$ (as returned by \algoname{ComputeMaxFlox}), which assigns to each
edge $(x_i,x_j)$ an integer $\phi(x_i,x_j)$ which represents the value of the flow on this edge. Here, we give a procedure
which produces an extension $P$ of $P_X$ where $x^*$ is a cowinner :
\begin{enumerate}
\item For each $x_i\in X_2$ and each $V\in \text{LP}(\widetilde{P},x_i)$, if there
is a flow from $x_i$ to $x_j$ via $V$, then we obtain $V^*$ from $V$
by putting two candidates from $Y$ in the top positions (that is,
both $\Delta(\widetilde{P},x_i)$ and $\Delta(\widetilde{P},x_j)$ are reduced by $1$, which comes at the cost
of using candidates in $Y$ twice). It is possible since $|Y|\geq
2$.
\item For each $x_i\in X_2$, if
$\phi(x_i,t)<\Delta(\widetilde{P},x_i)$, then we arbitrarily choose $\Delta(\widetilde{P},x_i)-\phi(x_i,t)$ votes
$V\in \text{LP}(\widetilde{P},x_i)$ among those which haven't been
selected in the previous step,
and obtain $V^*$ by putting two candidates from $Y$ in the top two
positions (again, we will specify how to choose the two candidates
from $Y$ later). It is possible since $|Y|\geq 2$.
\item For each $x_j\in X_1$, if $\phi(s,x_j)<\Delta(\widetilde{P},x_i)$, then we
arbitrarily choose $\Delta(\widetilde{P},x_i)-\phi(s,x_j)$ votes $V\in \text{HP}(\widetilde{P},x_j)$
such that $V^*$ is not defined above (in (1) or (2)), and then we
obtain $V^*$ by putting exactly one candidate from $Y$ in the top
position of $V$. This \jl{is possible because, by construction,}
$|\text{HP}(\widetilde{P},x_j)|=|\text{HP}(P,x_j)|\geq \Delta(P,x_i)\geq \Delta(\widetilde{P},x_i)$ for $x_j\in X_1$.
\item For each $V^*$, if a candidate $y \in Y$ is not selected for one of the first two
positions, then it is ranked at the bottom position.
\end{enumerate}
In the above procedure (similarly to what is done in Algorithm \ref{algo:addnew}), priority is given to candidates from $Y$ with the lowest score (or the lowest index, in case of ties) when it comes to choose those to be added on top of the votes.
Let us now determine the number of times that new candidates from $Y$ are inserted on top of the votes. Recall that
until line \ref{CC_build} of the algorithm, we have used the
candidates from $Y$ exactly $T$ times.
Now consider the four-step procedure described above.
Observe that to reduce by one unit the score deficit with respect to one candidate,
steps 1 and 3 require one occurrence of a candidate of $Y$ (step 1 uses two occurrences but reduces the score deficit with respect to two candidates), while step 2 requires \emph{two} occurrences.
Thus, for each $i\leq m-1$, we have to
use $\Delta(\widetilde{P},x_i)$ times the candidates from $Y$, plus the additional occurrences required in step 2.
More precisely, step 2 requires, for each $x_i \in X_2$, $\Delta(\widetilde{P},x_i) - \phi(x_i,t)$ additional occurrences of new candidates in the completed votes. Therefore, the total number of times that the
candidates of $Y$ are ranked either in first or second position (denoted $s_Y$ for readability), is such that:
\begin{align}
s_Y \leq & \sum_{i\leq m-1}\Delta(\widetilde{P},x_i)+(\sum_{x_i\in
X_2}\Delta(\widetilde{P},x_i)-\sum_{x_i\in X_2} \phi(x_i,t)) \\
= & \sum_{i\leq
m-1}\Delta(\widetilde{P},x_i)+(\sum_{x_i\in
X_2}\Delta(\widetilde{P},x_i)-\phi) \label{Sy_1}
\end{align}
But we also have :
\begin{align}
\phi \geq \sum_{i\leq
m-1}\Delta(\widetilde{P},x_i)+\sum_{x_i\in
X_2}\Delta(\widetilde{P},x_i)-(|Y| \cdot s(x^*,P_X)-T) \label{Sy_2}
\end{align}
By combining (\ref{Sy_1}) and (\ref{Sy_2}), we thus get :
\begin{align*}
s_Y \leq & |Y| \cdot s(x^*,P_X)-T \\
\leq & |Y| \cdot s(x^*,P_X)
\end{align*}
That is, our algorithm will put candidates from $Y$ in the top two positions in
the extension no more than $|Y|\cdot s(x^*,P_X)$ times. Because the addition of new candidates is done in a cyclic order, each new candidate will eventually appear at most $s(x^*,P_X)$ in the top two positions of the votes. Thus, the score of these new candidates will not exceed that of $x^*$. It follows that $x^*$ is a cowinner \jl{in} $P^*$, since for all other candidates $x_i \in X$, we have $\Delta(P^*,x_i)\le 0$.
Next, we show that if $x^*$ is a possible \jl{cowinner}, then the value of a max-flow must be at least
$$\sum_{i\leq
m-1}\Delta(\widetilde{P},x_i)+\sum_{x_i\in
X_2}\Delta(\widetilde{P},x_i)-(|Y| \cdot s(x^*,P_X)-T)$$
Due to
Proposition~\ref{prop:opt}, each extension profile $P^*$ of $P_X$
where $x^*$ becomes a cowinner to the problem instance can be
converted to a profile $\widetilde{P}$ as in the steps before line
\ref{CC_build} in the algorithm. Now, for each $x_i\in X_2$, let
$l_i$ denote the number of votes $V\in \text{LP}(\widetilde{P},x_i)$ such
that in its extension $V^*$, the top two positions are the
candidates of $Y$. We must have that $l_i\geq
\Delta(\widetilde{P},x_i)$. For every $x_i\in X_2$, we arbitrarily
choose $l_i-\Delta(\widetilde{P},x_i)$ such votes, and move the
first ranked candidate to the bottom position. For each $x_j\in
X_1$, let $l_j$ denote the number of votes $V\in
\text{HP}(\widetilde{P},x_j)\cup \text{MP}(\widetilde{P},x_j)$ such that in its
extension $V^*$, a candidate from $Y$ is ranked in the top
position. We must have that $l_j\geq \Delta(\widetilde{P},x_j)$. For
every $x_j\in X_1$, we arbitrarily choose
$l_j-\Delta(\widetilde{P},x_j)$ such votes, and move the first
ranked candidate to the bottom position.
Now, let there be a flow from $x_j \in X_1$ to $x_i \in X_2$ via
$V$ if $V\in \text{LP}(\widetilde{P},x_i)$ and the top two positions
in $V^*$ are both in $Y$. This defines a flow whose value is at
least $\sum_{x_i\in X_2}\Delta(\widetilde{P},x_i)-\sum_{x_j\in
X_1}(l_j-\Delta(\widetilde{P},x_j))$. Because the score of each
candidate of $Y$ is no more than $s(x^*,P_X)$, we know that $|Y|
\cdot s(x^*,P_X)-T\geq \sum_{i\leq m-1}l_i$. Actually, $|Y|
\cdot s(x^*,P_X)$ is the maximum score that the whole set of new
candidates of $Y$ can reach in such a way that $x^*$ is a cowinner.
In the partial profile $\widetilde{P}$ (line \ref{CC_build} of Algorithm
\algoname{CheckCowinner}$(P_X,x^{*},Y)$), the global score of $Y$ is $T$.
Finally, since $\sum_{i\leq m-1}l_i+T$ corresponds to the global
score that $Y$ has in profile $P^*$ (where $x^*$ becomes a
cowinner), we get $|Y| \cdot s(x^*,P_X)\geq \sum_{i\leq m-1}l_i+T$.
Hence, $|Y| \cdot s(x^*,P_X)-T\geq \sum_{i\leq m-1}l_i \geq
\sum_{x_i\in X_2}\Delta(\widetilde{P},x_i)+\sum_{x_j\in X_1}l_j$, or
equivalently, $-\sum_{x_j\in X_1}l_j\geq \sum_{x_i\in
X_2}\Delta(\widetilde{P},x_i)-(|Y| \cdot s(x^*,P_X)-T)$. Hence, we get:
\begin{align*}
\phi \geq &\sum_{x_i\in X_2}\Delta(\widetilde{P},x_i)-\sum_{x_j\in X_1}(l_j-\Delta(\widetilde{P},x_j))\\
=&\sum_{i\leq m-1}\Delta(\widetilde{P},x_i)-\sum_{x_j\in X_1}l_j\\
\geq &\sum_{i\leq m-1}\Delta(\widetilde{P},x_i)+\sum_{x_i\in
X_2}\Delta(\widetilde{P},x_i)-(|Y|s(x^*,P_X)-T)
\end{align*}
Thus, we have shown that $x^*$ is a possible cowinner if and only if the value of the flow from $s$ to $t$ is at least $\sum_{i\leq m-1}\Delta(\widetilde{P},x_i)+\sum_{x_i\in X_2}\Delta(\widetilde{P},x_i)-(|Y| \cdot s(x^*,P_X)-T)$. This concludes the proof.
\hfill \qed\\
\begin{coro}
\label{coro-2app}
Deciding whether $x^*$ is a possible cowinner for 2-approval with
respect to the addition of new candidates is in {\sf P}.
\end{coro}
To better understand Algorithm 1, we will now run it step by step on the example introduced previously.
\begin{example}
Consider the profile described in Example 1. We assume the number of new candidates is $k = 3$.
First, the initial scores of the candidates are $s(x^*,P_X) = 4$, $s(x_1,P_X) = 7$,
$s(x_2,P_X) = 7$, $s(x_3,P_X) = 8$, $s(x_4,P_X) = 6$ and
$s(x_5,P_X) = 4$ and $s(x_6,P_X) = 2$. The candidates
whose score exceeds that of
$x^*$ are $x_1$, $x_2$, $x_3$ and $x_4$, with the score
differences $\Delta(P,x_1) = 3$, $\Delta(P,x_2) = 3$, $\Delta(P,x_3)
= 4$ and $\Delta(P,x_4) = 2$. At first phase, we check if there are
candidates $x_i$ for which $|HP(P_X,x_i)| \leq \Delta(P_X,x_i)$.
This is the case for $x_1$, $x_2$ and $x_3$, thus we put one new
candidate on top of $v_8$, $v_9$, $v_{13}$ and $v_{14}$. The updated
table is as follows:
$$\begin{array}{|l|l|l|l|l|}
\hline
& \mbox{\text{\rm HP}} & \mbox{\text{\rm MP}} & \mbox{\text{\rm LP}} & \Delta(P,x_i)\\ \hline
x_1 & & v_1 & v_5, v_6, v_7 & 1 \\
x_2 & & v_8', v_9' & v_{10}, v_{11}, v_{12} & 2 \\
x_3 & & v_{13}', v_{14}' & v_{15}, v_{16}, v_{17}, v_{18} & 3 \\
x_4 & v_5, v_6, v_{10}, v_{15}, v_{16} && v_{19} & 2\\
\hline
\end{array}$$
Here, $v_i'$ refers to the vote $v_i$ to which new candidates have been added.
At the second phase, $X_2 = \{x_1,x_2,x_3\}$ and $X_1 = \{x_4\}$ (we do not worry about $x_5$ and $x_6$ for which nothing special has to be done).
We put one new candidate on top of $v_1$, $v_8'$, $v_9'$, $v_{13}'$ and $v_{14}'$, and we are done with $x_1$ and $x_2$
(since $\Delta(P,x_1)=0$ and $\Delta(P,x_2)=0$). The profile is now $\widetilde{P}$ and the updated table is :
$$\begin{array}{|l|l|l|l|l|}
\hline
& \mbox{\text{\rm HP}} & \mbox{\text{\rm MP}} & \mbox{\text{\rm LP}} & \Delta(\widetilde{P},x_i)\\ \hline
x_3 & & & v_{15}, v_{16}, v_{17}, v_{18} & 1\\
x_4 & v_5, v_6, v_{10}, v_{15}, v_{16} && v_{19} & 2 \\
\hline
\end{array}$$
So far we have used the new candidates 9 times, and
$s(x^*,\widetilde{P}) = 4$, therefore if we have less than three new
candidates we stop ($x^*$ is not a possible cowinner) otherwise we
continue. Now the situation is as follows and we have to solve the
corresponding maxflow problem (we omit the value of edges when it
equals 1).
\begin{footnotesize}
\[
\begin{array}{ccccccccccccccccccc}
v_1 & v_2 & v_3 & v_4 & v_5 & v_6 & v_7 & v_8 & v_9 & v_{10} & v_{11} & v_{12} & v_{13} & v_{14} & v_{15} & v_{16} & v_{17} & v_{18} & v_{19} \\
\hline
\bullet & x_1 & x_2 & x_3 & x_1 & x_1 & x_1 & \bullet & \bullet & x_2 & x_2 & x_2 & \bullet & \bullet & x_3 & x_3 & x_3 & x_3 & x_4 \\
x^* & x^* & x^* & x^* & x_4 & x_4 & x_5 & \bullet & \bullet & x_4 & x_5 & x_5 & \bullet & \bullet & x_4 & x_4 & x_5 & x_6& x_6 \\
\vdots & \vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots
\end{array}
\]
\end{footnotesize}
\begin{figure}[h!] \label{figureflow}
\begin{center}
\begin{tikzpicture}[scale=1]
\tikzstyle{s}=[circle,draw, line width=1pt, minimum size = 9mm, font= \footnotesize]
\node (s) at (0,2) [s] {$s$};
\node (x4) at (2,2) [s] {$x_4$};
\node (v15) at (4,1) [s] {$v_{15}$};
\node (v16) at (4,2) [s] {$v_{16}$};
\node (v17) at (4,3) [s] {$v_{17}$};
\node (v18) at (4,4) [s] {$v_{18}$};
\node (x3) at (6,2) [s] {$x_3$};
\node (t) at (8,2) [s] {$t$};
\draw [->] (s) -- (x4);
\draw [->] (x4) -- (v15);
\draw [->] (x4) -- (v16);
\draw [->] (v15) -- (x3);
\draw [->] (v16) -- (x3);
\draw [->] (v17) -- (x3);
\draw [->] (v18) -- (x3);
\draw [->] (x3) -- (t);
\end{tikzpicture}
\caption{The flow graph returned by \algoname{BuildMaxFlowGraph}$(\widetilde{P},x^*,\{x_4\},\{x_3\})$.} \label{flow}
\end{center}
\end{figure}
The maximum flow has value 1 and is obtained for instance by having a
flow 1 for instance through the edges $s \rightarrow x_4$, $x_4
\rightarrow v_{16}$, $v_{16} \rightarrow x_{3}$, $x_{3} \rightarrow
t$ (going through $v_{15}$ is an equally good option). Therefore we
place two new candidates on top of $v_{16}$, which has the effect of making
the score of $x_3$ and $x_4$ decrease by one unit each. We still
have to make the score of $x_4$ decrease by one unit, and for this
we must place one new candidates on top of any of the votes $v_5$, $v_6$,
$v_{10}$, $v_{15}$ (say $v_5$). In total we will have used the new
candidates 12 times, therefore, $c$ is a possible cowinner if and
only if the number of new candidates is at least 3. A possible
extension (with 3 new candidates) is as follows:
\begin{footnotesize}
\[
\begin{array}{ccccccccccccccccccc}
v_1 & v_2 & v_3 & v_4 & v_5 & v_6 & v_7 & v_8 & v_9 & v_{10} & v_{11} & v_{12} & v_{13} & v_{14} & v_{15} & v_{16} & v_{17} & v_{18} & v_{19} \\
\hline
{\bf y_1} & x_1 & x_2 & x_3 & {\bf y_2} & x_1 & x_1 & {\bf y_1} & {\bf y_2} & x_2 & x_2 & x_2 & {\bf y_1} & {\bf y_2} & x_3 & {\bf y_1} & x_3 & x_3 & x_4 \\
x^* & x^* & x^* & x^* & x_1 & x_4 & x_5 & {\bf y_3} & {\bf y_3} & x_4 & x_5 & x_5 & {\bf y_3} & {\bf y_3} & x_4 & {\bf y_2} & x_5 & x_6 & x_6 \\
\vdots & \vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots &\vdots
\end{array}
\]
\end{footnotesize}
\end{example}
\subsection{$K$-approval, two new candidates}\label{kapp-2}
Let $X = \{x^*\} \cup \{x_1,\ldots, x_p\}$ be the set of (initial) candidates, $x^*$ being the candidate that we want to make a cowinner,
$Y=\{y_1,y_2\}$ the two new candidates, and $P_X = \langle V_1,\ldots, V_n\rangle$ the initial profile,
where each $V_i$ is a sequence of $K$ candidates in $X$.
We first introduce the following notation:
\begin{itemize}
\item For each $x \in X$, $U^{P_X}(x)$ is the number of votes in $P_X$ whose candidates ranked $K-1$ and $K$ are respectively $x$ and $x^*$, and
$T^{P_X}(x) = S_{K-2}(x,P_X) + U^{P_X}(x)$. (Recall that
$S_{K-2}(x,P_X)$ is the number of voters in $P_X$ who rank $x$ in
the first $K-2$ positions.)
\end{itemize}
We establish the following lemma.
\begin{lemma}\label{lemma-2app}
For each $x \in X$, there exists a completion $Q$ of $P_X$ by adding
two candidates such that $S_K(x,Q) \leq S_K(x^*,Q)$ if and only if
$T^{P_X}(x) \leq S_K(x^*,P_X)$.
\end{lemma}
\noindent {\em Proof:~}
Assume $T^{P_X}(x) > S_K(x^*,P_X)$, and let $Q$ be a completion of
$P_X$ by adding two candidates \jl{in which $x^*$ is a cowinner}. Let us partition $P_X$ into $P_1$,
$P_2$ and $P_3$, as follows: every vote in $P_1$ is such that the
candidates ranked $K-1$ and $K$ are respectively $x$ and $x^*$;
$P_2$ contains all votes ranking $x$ in the first $K-2$ positions;
and $P_3$ contains all other votes in $P_X$. Let $Q_1$, $Q_2$ and
$Q_3$ be the corresponding votes in $Q$, and let $\alpha$ be the
number of votes in $Q_1$ where the two new candidates have been placed in
the first $K$ positions, thus eliminating both $x$
and $x^*$ from the $K$ first positions; clearly, we have $S_K(x,Q_1)
= S_K(x,P_1)-\alpha = U^{P_X}(x) - \alpha$ and $S_K(x^*,Q_1) \leq
S_K(x^*,P_1)-\alpha$ (the inequality can be strict, in case there
are some votes in $Q_1$ where only one new candidate was placed
in the first $K$ positions). Now, regardless of the position of
the two new candidates, we have $S_K(x,Q_2) = S_{K-2}(x,P_2)$. We
get $S_K(x,Q) = S_K(x,Q_1) + S_K(x,Q_2) + S_K(x,Q_3) \geq
U^{P_X}(x) - \alpha + S_{K-2}(x,P_2) = T^{P_X}(x) - \alpha$, whereas
$S_K(x^*,Q) \leq S_K(x^*,P_X) - \alpha$. The initial assumption
$T^{P_X}(x) > S_K(x^*,P_X)$ implies $T^{P_X}(x)- \alpha > S_K(x^*,P_X)
- \alpha$, therefore $S_K(x,Q) > S_K(x^*,Q)$.
Conversely, assume $T^{P_X}(x) \leq S_K(x^*,P_X)$, and let us build
$Q$ as follows: we introduce one new candidate on top of each vote
of $P_X$ that ranks $x$ in position $K$, and two new candidates on
top of each vote of $P_X$ that ranks $x$ in position $K-1$ and $x'
\neq x^*$ in position $K$. It is easy to check that $S_K(x^*,Q) =
S_K(x^*,P_X)$. Now, the only votes of $Q$ where $x$ remains among
the first $K$ position are those of $Q_1$ and of $Q_2$, therefore
$S_K(x,Q) = T^{P_X}(x) \leq S_K(x^*,P_X) = S_K(x^*,Q)$.
\hfill \qed\\
\begin{prop}\label{prop-2cand}
Deciding whether $x^*$ is a possible cowinner for $K$-approval with
respect to the addition of 2 new candidates is in {\sf P}.
\end{prop}
\noindent {\em Proof:~}
A consequence of Lemma \ref{lemma-2app} is that if $T^{P_X}(x) >
S_K(x^*,P_X)$ for some $x$, then $x^*$ cannot be a possible cowinner in $P_X$
under $2$-approval with 2 new candidates; and obviously, checking
whether $T^{P_X}(x) > S_K(x^*,P_X)$ holds for some $x$ can be done
in polynomial time. Therefore, {\em from now on, we assume that
$T^{P_X}(x) \leq S_K(x^*,P_X)$ holds for every $x \in X$} ---
assuming this will not change the complexity of the problem.
We now give a polynomial reduction from the possible cowinner
problem for $K$-approval and $2$ new candidates to the possible
cowinner problem for $2$-approval and $2$ new candidates, which we
already know to be polynomial. Let
$\langle N, X, P_X, 2\rangle$
be an instance of the possible cowinner problem for $K$-approval
with respect to the addition of $2$ new candidates.
We build an instance $\langle N', X', R_{X'}, 2\rangle$
of the possible cowinner problem for $2$-approval
with respect to the addition of $2$ new candidates in the
following way. The profile $P_X$ is translated into the following
profile $R=R_{X'}$:
\begin{itemize}
\item the set of candidates is $X'=X\cup \{z_j, 1 \leq j
\leq \sum_{x \in X \setminus \{x^*\}}S_{K-2}(x,P_X)\} \cup \{z_j', 1
\leq j \leq S_{K-2}(x^*,P_X)\}$, where all $z_j$ and $z_j'$ are
fresh candidates;
\item for every vote $V_i$ in $P_X$, we have in $R$ a vote $W_i$ including the candidates
ranked in positions $K-1$ and $K$ of $V_i$, and then the remaining candidates in any order. We denote by $R_1$ be the resulting set of votes;
\item for every $x \in X\setminus \{x^*\}$, we have $S_{K-2}(x,P_X)$ votes $x z_j$, and then the remaining candidates in any order.
We denote by $R_2$ the resulting set of votes;
\item similarly, we have $S_{K-2}(x^*,P_X)$ votes $z_j' x^*$, and then the remaining candidates in any order.
We denote by $R_3$ the resulting set of votes.
\end{itemize}
We note that if $x \in X$ then $S_{K}(x,P_X)=S_2(x,R)$, and for
every fresh candidate $z$, $S_2(z,R)=1$. Without loss of generality
we assume $S_K(x^*,P_X) \geq 1$ (otherwise we know for sure that
$x^*$ cannot be a possible cowinner).
We decompose the rest of the proof into two lemmas.
\begin{lemma}\label{lemmaprop5-1} If $x^*$ is a possible cowinner for $K$-approval with 2 new candidates in $P_X$ , then it is is a possible cowinner for 2-approval with 2 new candidates in $R$.
\end{lemma}
\noindent {\em Proof:~}
Suppose that $x^*$ is a possible cowinner for $K$-approval with 2
new candidates $Y=\{y_1,y_2\}$ in $P_X$ and let $P' = \langle V_1',
\ldots, V_n'\rangle$ be an extension of $P_X$ with two new
candidates where $x^*$ is a cowinner. Let us use these two new
candidates in the same way in $R$: every time a new candidate is
used for being placed on top of $V_i$, it is also used for being placed on top of $W_i$. Let
$R'$ be the resulting profile. All candidates in $X$ have the same
scores in $P_X$ and in $R$, they also will have the same scores in
$P'$ and $R'$; as for the fresh candidates $z_j,z'_j$, $S_2(z_j,R')
=S_2(z'_j,R')= 1 \leq S_2(x^*,R')$; therefore, $x^*$ is a cowinner
in $R'$ and a possible cowinner for 2-approval with 2 new candidates
in $R$.
\hfill \qed\\
\begin{lemma}\label{lemmaprop5-2} If $x^*$ is a possible cowinner for $K$-approval with 2 new candidates in $R$ , then it is a possible cowinner for 2-approval in $P_X$.
\end{lemma}
\noindent {\em Proof:~}
Suppose that $x^*$ is a possible cowinner for
$2$-approval with 2 new candidates $Y=\{y_1,y_2\}$ in $R$, and let
$R'$ be a completion of $R$ where $x^*$ is a cowinner for
$2$-approval. Let us write $R' = R_1' \cup R_2' \cup R_3'$, where
$R_1'$ (resp. $R_2'$, $R_3'$) consists in the completions of the
votes in $R_1$ (resp. $R_2$, $R_3$). By a slight abuse of language
we denote by $R_1, R_1'$ etc. only the part of the votes in
$R_1,R_1'$ etc. consisting of the top two candidates only.
We first claim that we can assume without loss of generality that
$R_2' = R_2$ and $R_3' = R_3$ that is, the only votes in $R'$ where
some new candidates have been placed on one of the top two positions
are in $R_1'$. Suppose this is not the case; then we are in one of
the following four situations: (1) there is a vote in $R_2'$ of the
form $y_j x_i$, where $y_j \in Y$ and $x_i \in X$, or (2) there is
a vote in $R_2'$ of the form $y_1 y_2$ or $y_2 y_1$, or (3) there
is a vote in $R_3'$ of the form $y_i z'_j$ or (4) there is a vote
in $R_3'$ of the form $y_1 y_2$ or $y_2 y_1$.
Consider first cases (1), (3) and (4). Take one of these votes in $R_2'$ (case (1)) or in $R_3'$ (cases (3) or (4))
and replace it by the original vote $x z_j$ in $R_2$ (case 1) or in $z'_j x^*$ in $R_3'$ (cases (3) or (4)). Let $R''$ be the profile obtained. We have
$S_2(x^*,R'') \geq S_2(x^*,R') \geq 1$,
for every $x_i \in X$, $S_2(x_i,R'') = S_2(x_i,R')$,
for every $z_j$, $S_2(z_j,R'') \leq 1$,
and for every $z_j'$, $S_2(z'_j,R'') \leq 1$.
Therefore, when transforming $R'$ into $R''$, the score of $x^*$
does not decrease whereas the score of all other candidates does not
increase; because $x^*$ is a cowinner in $R'$, it is still a
cowinner in $R''$. Lastly, $R''$ is also an extension of $R$.
By induction, if we perform this operation for each occurrence of
cases (1), (3) or (4), we end up with a profile $R''$, which is an
extension of $R$ for which situations (1), (3) and (4) do not
occur, and such that $x^*$ is a cowinner for 2-approval in $R''$.
Let $R'' = R_1'' \cup R_2'' \cup R''_3 = R_1'' \cup R_2'' \cup R_3$.
Now, consider case (2). Let $x_i z_j$ be one of the votes in $R_2$
corresponding to a vote $y_1 y_2$ (or $y_2 y_1$) in $R_2''$. Apply
the following procedure in this order:
\begin{enumerate}
\item Assume that $x_i$ does not appear in $R_1''$ except in votes of the form $x_i x^*$, and let $R'''$
be the profile obtained from $R''$ by replacing the vote $y_1 y_2$
in $R_2''$ by the original vote $x_i z_j$ in $R_2$. Then
$S_2(x_i,R''') = S_2(x_i,R_1''') + S_2(x_i,R_2''') \leq
S_2(x_i,R_1''') + S_2(x_i,R_2)$. Now, $S_2(x_i,R''') \leq S_2(x_i,R''') = U^{P_X}(x_i)$ and $S_2(x_i,R_2) = S_{K-2}(x_i,P_X)$, therefore $S_2(x_i,R''') \leq$ $U^{P_X}(x_i) + S_{K-2}(x_i,P_X) =
T^{P_X}(x_i) \leq S_K(x^*,P_X) = S_2(x^*,R) = S_2(x^*,R''')$.
Therefore, $x^*$ is also a cowinner in $R'''$.
\item Now, assume that $x_i$ appears in at least one vote of $R_1''$ of the form $x^* x_i$, $x_i x_j$ or $x_j x_i$.
If this is a vote of the form $x_i x_j$, we replace $y_1 y_2$ in
$R_2''$ by the original vote $x_i z_j$ in $R_2$ and the vote $x_i
x_j$ by a vote $y_1 y_2$. If this is a vote of the form $x^* x_i$ or
$x_j x_i$, we replace $y_1 y_2$ in $R_2''$ by the original vote $x_i
z_j$ in $R_2$ and the vote $x^* x_i$ (resp. $x_j x_i$) by $y_1 x^*$
(resp. $y_1 x_j$). In all three cases, the score of all candidates
remain the same after the transformation, except the score of $y_2$
and $z_j$, which can only decrease, therefore $x^*$ is still a
cowinner after the transformation.
\end{enumerate}
We perform this procedure on $x_i$ iteratively until all the votes $y_1 y_2$ (or $y_2 y_1$) in $R_2'''$
have been replaced by the original votes $x_i z_j$ in $R_2$. After doing this sequentially on all candidates
of $X$ such that case (2) occurs, we end up with a profile $R''''$ of $P$ such that $R_2'''' = R_2$ and $R_3'''' = R_3$ and $x^*$ is a cowinner in $R''''$.
This proves the claim.\medskip
Now, let $R'$ be a completion of $R$ where $x^*$ is a cowinner for $2$-approval, where $R_2' = R_2$ and $R_3' = R_3$.
From $R'$ we build the following extension $P'$ of $P$: for every vote $W_i \in R_1$,
\begin{itemize}
\item if $W_i'$ is of the form $x^* x'$ then $V_i' = W_i'$;
\item if $W_i'$ is of the form $y_i x$ then $V_i'$ is obtained from $V_i$ by placing $y_i$ on top;
\item if $W_i'$ is of the form $y_1 y_2$ (or $y_2 y_1$) then $V_i'$ is obtained from $V_i$ by placing $\{y_1, y_2\}$ on top.
\end{itemize}
The scores of all candidates are the same in $P'$ and in $R'$, therefore $x^*$ is a cowinner in $P'$ if only if it is a cowinner in $R'$.
Therefore, it is a cowinner in $P'$, which means that $x^*$ is a possible cowinner for 2-approval in $P$.
\hfill \qed\\
We can now end the proof of Proposition \ref{prop-2cand}: from Lemmas \ref{lemmaprop5-1} and \ref{lemmaprop5-2} we conclude that deciding whether $x^*$ is a possible cowinner for $K$-approval with respect to the addition of two candidates can be polynomially reduced to a problem of a deciding whether $x^*$ is a possible cowinner for $2$-approval, which we know is in {\sf P}.
\hfill \qed\\
\subsection{$3$-approval, $3$ new candidates}\label{kapp-2}
We will now see that the problems addressed in previous subsections constitute the frontier of what can be solved in polynomial-time for $K$-approval rules.
In the rest of this paper, the hardness proofs will use reductions from the 3-dimensional matching (\textsc{3-DM}) problem.
\begin{defin}\label{def:3DM}
An instance of \textsc{3-DM} consists of a subset
$\mathcal{C}=\{e_1,\dots,e_m\}\subseteq A\times B\times C$ of
triples, where $A,B,C$ are 3 pairwise disjoint sets of size $n'$ with
$A=\{a_1,\dots,a_{n'}\}$, $B=\{b_1,\dots,b_{n'}\}$ and
$C=\{c_1,\dots,c_{n'}\}$.
For $z\in A\cup B\cup C$, $d(z)$
denotes the number of occurrences of $z$ in $\mathcal{C}$, that is
the number of triples of $\mathcal{C}$ which contain $z$.
A matching is a subset
$M\subseteq\mathcal{C}$ such that no two elements in $M$ agree on
any coordinate.
The \textsc{3-DM} problem consists in answering this question: does there
exist a perfect matching $M$ on $\mathcal{C}$, that is, a matching
of size $n'$?
\end{defin}
The \textsc{3-DM} problem is known to be
{\sf NP}-complete (problem [SP1] page 221 in \cite{GJ79}),
even with the restriction where $\forall z\in A\cup B\cup C$, $d(z)\in\{2,3\}$ (that is, no element of
$A\cup B\cup C$ occurs in more than 3 triples, and each element of
$A\cup B\cup C$ appears in at least 2 triples).
\begin{prop}\label{PropNP-completeApproval}
Deciding if $x^*$ is a possible cowinner for 3-approval with respect
to the addition of 3 new candidates, is an {\sf NP}-complete problem.
\end{prop}
\noindent {\em Proof:~}
This problem is clearly in {\sf NP}. The hardness proof is based on a reduction
from \textsc{3-DM} (see Definition \ref{def:3DM}).
Let $I=(\mathcal{C},A\times B\times C)$ be an instance of
\textsc{3-DM} with $n'\geq 3$ and $\forall z\in A\cup B\cup C$, $d(z)\in\{2,3\}$.
From
$I$, we build an instance of the PcWNC problem as follows. The set
$X$ of candidates contains $x^*$, $X_1= \{x'_i,y'_i,z'_i:1\leq i\leq
n'\}$ where $x'_i,y'_i,z'_i$ correspond to elements of $A\cup B\cup
C$ and a set $X_2$ of dummy candidates. We now describe the votes
informally; their formal definition will follow.
The set $N$ of voters
contains $N_1=\{v^e:e\in\mathcal{C}\}$ and a set $N_2$ of dummy
voters. For each voter, we only indicate her \jl{first three} candidates.
Thus, the vote of $v^e$ is $(x'_i,y'_j,z'_k)$ where
$e=(a_i,b_j,c_k)\in\mathcal{C}$. The preference of dummy voters are
such that :
\begin{itemize}
\item [$(i)$] the \jl{scores of the candidates in $X$ satisfy} $\forall
x\in X_1$, $S_{3}(x,P_X)=n'+1$, $S_{3}(x^*,P_X)=n'$ and $\forall
x\in X_2$, $S_{3}(x,P_X)=1$;
\item[$(ii)$] the vote of any voter of $N_2$ contains
at most one candidate from $\{x'_i,y'_i,z'_i:1\leq i\leq n'\}$ in
the first three positions, and if it contains one, then it is in top
position.
\end{itemize}
Formally, the instance $\langle N, X, P_X, 3, x^*\rangle$
of the possible cowinner problem for $3$-approval and $3$ new candidates
is described as follows: the set
of voters is $N=N_1\cup N_2$ where $N_1=\{v^e:e\in\mathcal{C}\}$ and
$N_2=N_A\cup N_B\cup N_C\cup N_{x^*}$, the set of candidates is
$X=X_1\cup X_2$. For the candidates in
$X$, we have $x^*$ together with :
\begin{itemize}
\item $X_1=X'\cup
Y'\cup Z'$ where $X'=\{x'_1,\dots,x'_{n'}\}$, $Y'=\{y'_1,\dots,y'_{n'}\}$
and $Z'=\{z'_1,\dots,z'_{n'}\}$.
\item $X_{2}=
\{x^*_i: 1\leq i\leq 2n'\}\cup \{x_i^j:1\leq i\leq n'$, $1\leq j\leq 2\left(n'-d(x_i)+1\right)\}\cup
\{y_i^j:1\leq i\leq n'$, $1\leq j\leq 2\left(n'-d(y_i)+1\right)\}\cup
\{z_i^j:1\leq i\leq n'$, $1\leq j\leq 2\left(n'-d(z_i)+1\right)\}$.
Note that $n'-d(x_i)+1\geq 1$ since $d(z)\leq 3\leq n'$.
\end{itemize}
For each voter $v_i\in N$, we only indicate her \jl{first three} candidates
(in the order of preference). The set of all $X$-votes ${\cal P}_X$ of the voters in $N$ is
as follows :
\begin{itemize}
\item $N_A=\{v_{i,j}^A:1\leq i\leq n'$,
$0\leq j \leq \left(n'-d(x_i)\right)\}$. The vote of $v_{i,j}^A$ is
$(x'_i,x_i^{2j+1},x_i^{2j+2})$.
\item $N_B=\{v_{i,j}^B:1\leq i\leq n'$,
$0\leq j \leq \left(n'-d(y_i)\right)\}$. The vote of $v_{i,j}^B$ is
$(y'_i,y_i^{2j+1},y_i^{2j+2})$.
\item $N_C=\{v_{i,j}^C:1\leq i\leq n'$,
$0\leq j \leq \left(n'-d(z_i)\right)\}$. The vote of $v_{i,j}^C$ is
$(z'_i,z_i^{2j+1},z_i^{2j+2})$.
\item $N_1=\{v^e:e\in\mathcal{C}\}$. The vote of
$v^e$ is $(x'_i,y'_j,z'_k)$ where $e=(a_i,b_j,c_k)\in\mathcal{C}$.
\item $N_{x^*}=\{v_j^{x^*}:0\leq j \leq n'-1\}$. The vote of
$v_j^{x^*}$ is $(x^*,x^*_{2j+1},x^*_{2j+2})$.
\end{itemize}
We claim that $I$ admits a perfect matching $M\subseteq \mathcal{C}$
if and only if $x^*$ becomes a possible cowinner by adding three new
candidates.
Let $Y=\{y_1,y_2,y_3\}$ be the new candidates added. Since we cannot
increase the score of $x^*$, we must decrease by one point the scores
of candidates of $X'\cup Y' \cup Z'$. Let us focus on candidates in
$X'$. In order to reduce the score of $x'_i$, we must modify the
votes of voters in $N_1$ or in $N_A$. By construction, each such
voter must put $y_1,y_2,y_3$ in the first three positions (since in $N_A$ or
from $(ii)$, candidates of $X'$ are put in top position when they
appear in the first three positions) and then, the score of each $y_i$ increases
by 1 at each time. Since there are $n'$ candidates in $X'$, we deduce
$S_{3}(y_i,P)\geq n'$ for every $i=1,2,3$. On the other hand, if
$x^*$ becomes a cowinner, $S_{3}(y_i,P)\leq S_{3}(x^*,P)\leq
S_{3}(x^*,P_X)=n'$ from $(i)$. Thus, $S_{3}(y_i,P)= n'$ for every
$i=1,2,3$ and there are exactly $n'$ voters $N'$ which put
$y_1,y_2,y_3$ in the first three positions (for the remaining voters of
$N\setminus N'$, $y_i$ is ranked
in position at least 4 for every $i=1,2,3)$.\\
We claim that $N'\subseteq N_1$. Otherwise, at least one voter of
$N_A$ put $y_1,y_2,y_3$ in the first three positions. There remains at
most $n'-1$ voters of $N'$ to decrease by 1 the score of candidates
in $Y'$. It is impossible because $|Y'|=n'$ and, from $(ii)$ and by
construction of $N_1$, each candidate of $Y'$ appears at most once
in the first three positions for all voters. Finally, since the score of
candidates in $Y'\cup Z'$ must also decrease by 1, we deduce that
$x^*$ is a possible cowinner iff $M=\{e\in \mathcal{C}:$
$y_1,y_2,y_3$ are in the first three positions for voter $v_e\}$ is a
perfect matching of $\mathcal{C}$.
\hfill \qed\\
\subsection{General case}
We finalize the study of the possible cowinner problem for $K$-approval with respect to candidate addition by showing that the problem is hard in any other case. For this we proceed in two steps: we first prove that for each $k\geq 3$,
the problem {\sc PcWNC}($k$)
for $3$-approval
is {\sf NP}-complete\ (Lemma \ref{Kappk>3}). Next we prove that if the problem {\sc PcWNC}($k$)
for $K$-approval
is {\sf NP}-complete\, then it is also the case for the problem {\sc PcWNC}($k$)
for $(K+1)$-approval (Lemma \ref{KappK+1}).
\begin{lemma}\label{Kappk>3}
For all $k \geq 3$, the problem {\sc PcWNC}$(k)$ for $3$-approval can be reduced in polynomial-time to the
problem {\sc PcWNC}$(k+1)$ for $3$-approval.
\end{lemma}
\noindent {\em Proof:~}
Let $\langle N, X,P_X, k, x^*\rangle$, where $P =P_X= \langle V_1, \ldots, V_n\rangle$,
be an instance of $\text{\sc PcWNC}(k)$
for $3$-approval.
Assume $S_3(x^*,P)\geq 1$ (otherwise, the problem is trivial). Consider the following instance
$\langle N',X',Q_{X'}, k+1, x^*\rangle$
of the $\text{\sc PcWNC}(k+1)$ for $3$-approval:
\begin{itemize}
\item the set of candidates is $X' = X \cup \{z\} \cup \{t_i^1,t_i^2 \ | \ 1 \leq i \leq 2S_3(x^*,P)\}$;
\item there are $n+2S_3(x^*,P)$ votes:
\begin{itemize}
\item for every vote $V_j$ in $P$ we have a vote $W_j$ in $Q$ whose first three candidates are
the same as in $V_j$ and in the same order, and the other candidates are in an arbitrary order.
\item for every $i = 1, \ldots, 2 S_3(x^*,P)$, we have a vote $U_i$ in which the first 3 candidates
are $t_i^1, t_i^{2}, z$, the remaining candidates being ranked arbitrarily.
\end{itemize}
\end{itemize}
Assume $x^*$ is a possible cowinner for $P=P_{X}$ (w.r.t. the
addition of $k$ new candidates) and let $P'$ be an extension of $P$
where $x^*$ is a cowinner. Let $Y=\{y_1,\ldots, y_k\}$ denote the
new candidates for the instance $\langle N,X, P_X, k\rangle$, and
$Y'=\{y_1,\ldots, y_{k+1}\}$ the new candidates for the instance
$\langle N', X', Q_{X'}, k+1\rangle$. Consider the following
extension $Q'$ of $Q=Q_{X'}$: for every vote $V'_j$ of $P'$ we have
a vote $W'_j$ in $Q'$ whose 3 first candidates are the same as in
$V'_j$ (and the remaining ones in an arbitrary order); and for every
vote $U_i$ such that $1 \leq i \leq S_3(x^*,P)$ we have a vote
$U'_i$ whose first 3 candidates are $y_{k+1}, t_i^1, t_i^{2}$ and
for every vote $U_i$ such that $S_3(x^*,P)+1 \leq i \leq 2 S_3(x^*,P)$, we have a vote
$U'_i$ whose first 3 candidates are $t_i^1, t_i^{2}, z$. It is easy
to check that $Q'$ is an extension of $Q$. The scores of all
candidates in $X \cup Y$ are the same in $P'$ and $Q'$, while the
score of each $t_i^1,t_i^2$ is 1, the scores of $z$ and of $y_{k+1}$
are $S_3(x^*,P)$; therefore $x^*$ is a cowinner in $Q'$ and a
possible cowinner in $Q$.
Conversely, assume $x^*$ is a possible cowinner in $Q=Q_{X'}$ and
let $Q'$ be an extension of $Q$ in which $x^*$ is a cowinner.
We are now going to reason abut the number of occurrences
of the new candidates $y_1,\ldots, y_{k+1}$ in the first three positions of the votes of $Q'$.
For the sake of notation, for any vote $V$ we denote $S_3(Y',V) = \sum_{y \in Y'}S_3(y',V)$:
in words, $S_3(Y',V)$ is the number of new candidates in the first three positions of $V$.
Similarly, if $R$ is a profile, we denote $S_3(Y',R)= \sum_{y \in Y'}S_3(y',R)$.
Without loss of generality, we assume that $S_3(x^*,Q') = S_3(x^*,Q)
= S_3(x^*,P)$, since under $3$-approval it is never beneficial to
decrease the score of $x^*$ to make it a possible cowinner.
We have $S_3(z,Q') \leq S_3(x^*,Q') = S_3(x^*,P)$ and $S_3(z,Q) = 2
S_3(x^*,P)$, therefore a new candidate must be put above $z$ in at
least $S_3(x^*,P)$ votes $U'_i$; therefore,
$$\sum_{j=1}^{2S_3(x^*,P)}S_3(Y',U_i) \geq S_3(x^*,P) ~~~(1)$$
Now, $S_3(Y',Q') = \sum_{i=1}^n S_3(Y',W_i) +\sum_{j=1}^{2S_3(x^*,P)}S_3(Y',U_i)$, which together with (1) entails
$$\sum_{i=1}^n S_3(Y',W_i) \leq S_3(Y',Q') - S_3(x^*,P) ~~~(2)$$
Now, $x^*$ is a cowinner in $Q'$, therefore, for all $y_j \in Y'$ we have $S_3(y_j,Q') \leq S_3(x^*,Q') = S_3(x^*,P)$, from which we get
$$S_3(Y',Q') \leq (k+1)S_3(x^*,P) ~~~(3)$$
From (2) and (3) we get
$$\sum_{i=1}^n S_3(Y',W_i) \leq kS_3(x^*,P) ~~~(4)$$
Now, consider the extension $P'$ of $P$ built from the restriction of $Q'$ to $\{W_1', \ldots, W_n'\}$ by changing the candidates in $Y$ placed in the first three positions {\em in such a way that each candidate appears at most in $S_3(x^*,P)$ votes}, which is made possible by (4). We have:
\begin{itemize}
\item $S_3(x^*,P') = S_3(x^*,P')$;
\item for each $y \in Y$, $S_3(y,P') \leq S_3(x^*,P) = S_3(x^*,P')$;
\item for each $x \in X \setminus \{x^*\}$, $S_3(x,P') = S_3(x,Q')$; because $x^*$ is a possible cowinner in $Q'$, we have $S_3(x,Q') \leq S_3(x^*,Q') = S_3(x^*,P)$, therefore,
$S_3(x,P') \leq S_3(x^*,P) = S_3(x^*,P')$.
\end{itemize}
From this we conclude that $x^*$ is a possible cowinner in $P'$.
\hfill \qed\\
\begin{lemma}\label{KappK+1}
The problem $\text{\sc PcWNC}(k+1)$ for $K$-approval
can be reduced in polynomial-time to the problem $\text{\sc PcWNC}(k)$ for $(K+1)$-approval.
\end{lemma}
\noindent {\em Proof:~}
Let $\langle N, X,P_X,
k, x^*\rangle$ where $P_X = \langle V_1, \ldots, V_n\rangle$
be an instance of PcWNC($k$)
for $K$-approval.
Consider the following instance $\langle N',X',R_{X'},k, x^*\rangle$
of the PcWNC($k$) for $(K+1)$-approval:
\begin{itemize}
\item the set of candidates is $X' = X \cup \{t_i \ | \ 1 \leq i \leq n\}$;
\item for every vote $V_i$ in $P$ we have a vote $W_i$ in $R$ whose top candidate is $t_i$
and the candidates ranked in position 2 to $K+1$ are the candidates ranked in positions 1 to $K$ in $V_i$,
the remaining candidates being ranked arbitrarily.
\end{itemize}
Assume $x^*$ is a possible cowinner for $P=P_X$ and let $P' =
\langle V_1', \ldots, V_n'\rangle$ be an extension of $P$ where
$x^*$ is a cowinner. Denote by $y_1, \ldots, y_k$ the new
candidates. Consider the extension $R' = \langle W_1', \ldots,
W_n'\rangle$ of $R=R_{X'}$ where $W_i'$ ranks $t_i$ first and then
the candidates ranked in the first $K$ positions in $V_i'$. For
every $x \in X$ we have $S_{K+1}(x,R') = S_{K}(x,P')$; for every $i
= 1,\ldots, k$ we have $S_{K+1}(y_i,R') = S_{K}(y_i,P')$; and for
every $j = 1, \ldots, n$, we have $S_{K+1}(t_j,R') = 1$. Therefore
$x^*$ is a possible cowinner in $R'$ and a possible cowinner in
$R$.
Conversely, assume $x^*$ is a possible cowinner in $R=R_{X'}$ and
let $R' = \langle W_1', \ldots, W_n'\rangle$ be a completion of $R$
in which it is a possible cowinner. Since none of the $t_i$
threatens $x^*$, without loss of generality we assume $t_i$ still
appears in the first $K+1$ positions of $W_i'$---otherwise,
change $W_i'$ by moving $t_i$ to the top of
$W_i'$. Consider now the extension $P' = \langle
V_1', \ldots, V_n'\rangle$ of $P=P_X$ where $V_i'$ is obtained from
$W_i'$ by removing all the $t$'s. Since exactly one $t_i$ appears in
the first $K+1$ positions of $W_i'$, the $K$ candidates approved in
$V_i'$ are exactly the $K+1$ candidates approved in $W_i'$ minus
$t_i$. From this we conclude that for every $x \in X$ we have
$S_{K+1}(x,P') = S_{K}(x,R')$ and for every $i = 1,\ldots, k$ we
have $S_{K+1}(y_i,P') = S_{K}(y_i,R')$.
Therefore $x^*$ is a possible cowinner in $P'$ and a possible cowinner in $P$.
\hfill \qed\\
\begin{prop}\label{44}
Deciding whether a candidate is a possible cowinner for $K$-approval with respect to the addition of
$k$ new candidates is {\sf NP}-complete\ for each $(K,k)$ such that $K \geq 3$ and $k \geq 3$.
\end{prop}
\noindent {\em Proof:~}
Since deciding whether $x^*$ is a possible cowinner for $3$-approval
with respect to the addition of $3$ new candidates is {\sf NP}-hard, using
inductively the reductions of Lemma \ref{Kappk>3} and Lemma \ref{KappK+1} shows that {\sf NP}-hardness
propagates to every $(K, k) \geq (3, 3)$. Hence, the problem
PcWNC($k$) for $K$-approval is
{\sf NP}-complete\ for any fixed pair of values $K \geq 3$ and $k \geq 3$.
\hfill \qed\\
We summarize the results obtained in this Section by the following table:
\begin{center}
\begin{tabular}{c|ccc}
& \nicobf{$k=1$} &\nicobf{$k=2$}& \nicobf{$k \geq 3$} \\ \hline
plurality & {\sf P} (Prop. \ref{prop-plura}) & {\sf P} (Prop. \ref{prop-plura}) & {\sf P} (Prop. \ref{prop-plura}) \\
2-approval & {\sf P} (Prop. \ref{prop-kapp}) & {\sf P} (Coro. \ref{coro-2app}) & {\sf P} (Coro. \ref{coro-2app})\\
$K$-approval, $K \geq 3$ & {\sf P} (Prop. \ref{prop-kapp}) & {\sf P} (Prop. \ref{prop-2cand}) & {\sf NP}-complete ~(Prop. \ref{44})
\end{tabular}
\end{center}
Observe that it would also be possible to address the {\sc PcWNC}$(k)$ problem (for $K \geq 3$ and $k \geq 3$) by
working out a direct polynomial reduction from 3-DM, as done in
Proposition \ref{PropNP-completeApproval}. This would however result
in a much less readable proof. One further interest of the proposed
reduction is to show how it is possible to ``neutralize'' the
(extended) power induced by adding more candidates by also adding
one more (dummy) candidate in the initial instance. Intuitively, by
setting the score of dummy candidate $t$ to $2 S_K(x^*,P)$, a single
new candidate $y_i$ will have to be ``consumed'' to ensure that $t$
does not win. More generally, the same proof holds even if $K$ and
$k$ depend on the instance (i.e. are not constant).
If we allow $f(n)$ new
candidates (where $f$ is polynomially bounded function) instead of
$k$ a constant, the hardness result also holds (in the proof of
Lemma \ref{Kappk>3}, we duplicate each vote $V$ $f(n)$ times by
adding candidates $z_i$ for $i=1,\dots,f(n)$ instead of $z$ and we
add dummy voters and candidates). Formally, we replace the
construction given in Lemma \ref{Kappk>3} by:
\begin{itemize}
\item the set of candidates is $X' = X \cup \{z_1,\ldots,z_{f(n)}\} \cup \{t_{i,\ell}^1,t_{i,\ell}^2 \ |
\ 1 \leq i \leq 2S_3(x^*,P), 1 \leq \ell \leq f(n) \}$;
\item there are $n+2f(n)S_3(x^*,P)$ votes:
\begin{itemize}
\item for every vote $V_j$ in $P$ we have a vote $W_j$ in $Q$ whose first three candidates are
the same as in $V_j$ and in the same order, and the other candidates are in an arbitrary order.
\item for every $i = 1, \ldots, 2 S_3(x^*,P)$ and $\ell = 1, \ldots, f(n)$,
we have a vote $U_{i,\ell}$ in which the first 3 candidates
are $t_{i,\ell}^1, t_{i,\ell}^{2}, z_\ell$, the remaining candidates being ranked arbitrarily.
\end{itemize}
\end{itemize}
Finally, $Y'=\{y_1,\ldots, y_{k+f(n)}\}$ are the new candidates.
Thus, using above construction, Lemma \ref{Kappk>3} and Proposition
\ref{PropNP-completeApproval}, we obtain that for any $\varepsilon
\in (0;1)$, {\sc PcWNC}$(f(n))$ for 3-approval is an {\sf NP}-complete
problem where
$f(n)=\Theta(|N|^{1-\varepsilon})=\Theta(|X|^{1-\varepsilon})$ (by
setting $f(n)=|N|^r$ in the above construction where $r$ is a
constant arbitrarily large). On the other hand, {\sc PcWNC}$(f(n))$ for
$K$-approval is a problem which can be solved in polynomial time when $f(n)=K \cdot |N|$, i.e., when the
number of new candidates is $K$ times the number of voters.
Note that some candidates
(other than the new candidates) can be \emph{necessary} cowinners with
$K$-approval. Specifically, each candidate $x_i$ such that
$S_{K-k}(P_X,x_i)=n$ is a necessary cowinner, since she is approved
by all voters and there are not enough new candidates to push her
(in at least one vote) out of the set of approved candidates.
\section{Borda}\label{borda}
Let us now consider the Borda rule ($r_B$).
Characterizing possible Borda cowinners when adding candidates is easy
due to the following lemma:
\begin{lemma}\label{lemma-borda}
Let $P_{X}$ be an $X$-profile where $X=\{x^*\}\cup
\{x_1,\ldots,x_p\}$ and let $Y=\{y_{1},\ldots, y_{k}\}$ be a set of $k$
new candidates. Let $r_{\vec{s}}$ be a scoring rule for $p+k$ candidates\footnote{In this lemma we do not have to deal with profiles with less than $p+k$ candidates, therefore it is not necessary to mention how $r_s$ is derived for fewer candidates than $p+k$.}
defined by the vector $\vec{s} = \left\langle s_{1}, \ldots , s_{p}, \ldots ,
s_{p+k}\right\rangle $ such that
$(s_{i}-s_{i+1})\le(s_{i+1}-s_{i+2})$ for all $i$. $x^*\in X$ is a
possible cowinner for $P_{X}$ w.r.t. the addition of $k$ new
candidates for the scoring rule $r_{\vec{s}}(P)$ iff $x^* \in r(P)$
where $P$ is the profile on $X\cup Y $ obtained from $P_{X}$ by
putting $y_{1}, \ldots , y_{k}$ right below $x^*$ (in arbitrary order)
in every vote of $P_{X}$.\end{lemma}
\noindent {\em Proof:~}
We show that it is never strictly better to put the new candidates
anywhere but right below $x$ in the new profile.
Let $P$ be an extension of $P_X$ in which $x^*$ is a cowinner, and
assume there is a vote $V\in P$ and a new candidate
$y$ such that either (i) $y \succ_v x^*$ or (ii) there exists at least one
candidate $x'$ such that $x^* \succ_v x' \succ_v y$.
If we are in case (i), let us move $y$ right below $x^*$; let $V'$
be the resulting vote, and $P'$ the resulting profile. Obviously,
$S_{\vec{s}}(y,P') \leq S_{\vec{s}}(y,P)$ and $S_{\vec{s}}(x^*,P')
\geq S_{\vec{s}}(x^*,P)$, therefore $S_{\vec{s}}(x^*,P') \geq
S_{\vec{s}}(y,P')$. For each candidate $z$ such that $y \succ_v z
\succ_v x^*$, let $i$ be the rank of $z$ in $v$ and $j > i$ be the
rank of $x^*$ in $v$. Then $(S_{\vec{s}}(z,P') -
S_{\vec{s}}(x^*,P')) - (S_{\vec{s}}(z,P)- S_{\vec{s}}(x^*,P)) =
(s_{i-1}-s_{j-1}) - (s_{i}-s_j) = (s_{i-1}-s_i) - (s_{j-1}-s_j) \leq
0$, therefore $S_{\vec{s}}(x^*,P') \geq S_{\vec{s}}(z,P') $. The
scores of all other candidates are left unchanged, therefore $x^*$ is
still a cowinner in $P'$. By applying this process iteratively for all new candidates and in
all votes until (i) no longer holds, we obtain a profile $Q$
in which $x^*$ is a cowinner, and such that $x^*$ is ranked above all new candidates in every vote.
Now, if (ii) holds for some new candidate $y$ and some vote $V$ of $Q$, then we move $y$ upwards,
right below $x^*$; let $V'$ be the resulting vote and $Q'$ the resulting
profile. The score of $y$ improves, but since $y$ is still ranked above all new candidates in every vote of $Q'$,
we have $S_{\vec{s}}(x^*,Q') \geq S_{\vec{s}}(y,Q')$. For each candidate $z \in X \cup Y \setminus \{x^*,y\}$
such that $x^* \succ_v z \succ_v y$ in vote $V$, $z$ moves down one
position in $Q'$, therefore $S_{\vec{s}}(z,Q') \leq S_{\vec{s}}(x',Q)
\leq S_{\vec{s}}(x^*,Q) =S_{\vec{s}}(x^*,Q')$. The scores of all
other candidates do not change, therefore $x^*$ is still a cowinner
in $Q'$. By applying this process iteratively and in all votes, until
(ii) no longer holds, and we obtain a profile in which
$x^*$ is a cowinner and neither (i) nor (ii) holds.
We conclude that $x^*$ is a possible cowinner for a profile if and only if
it is a cowinner in an extension of the profile where all new
candidates have been placed right below $x^*$.
\hfill \qed\\
In words, Lemma \ref{lemma-borda} applies to the rules
where the difference of scores between successive ranks can only
become smaller or remain constant as we come closer to the highest
ranks. This condition is satisfied by Borda (but not by plurality),
by veto, and by rules such as ``lexicographic veto'', where the
scoring vector is $ \langle M^p, M^p-M, M^p - M^2, \ldots, M^p -
M^{p-1}, 0\rangle$ where $M > n$.
The following result then easily follows:
\begin{prop}\label{prop-borda}
Let $P_{X}$ be an $X$-profile where $X=\{x^*\}\cup
\{x_1,\ldots,x_p\}$ and let $Y=\{y_{1},\ldots, y_{k}\}$ be a set of $k$
new candidates. A candidate $x^*$ is a possible cowinner for Borda
with respect to the addition of $k$ new candidates if and only if
$$k \geq \max_{z \in X \setminus \{x^*\}}\frac{S_{B}(z,P_X) - S_{B}(x^*,P_X)}{N_{P_X} (x^*,z)}$$
\end{prop}
\noindent {\em Proof:~}
By Lemma \ref{lemma-borda}, $x^*$ a possible cowinner if and only if it is a cowinner in the $X \cup \{y_1 , \dots , y_k\}$-completion $P$ of $P_X$ where $y_1, \dots , y_k$ are placed right below $x^*$, that is, if and only if
$S_{B}(x^*,P)=S_{B}(x^*,P_{X})+ kn$. Now, for each vote, all candidates in $X \setminus \{x^*\}$
ranked above $x^*$ get $k$ additional points in the extended vote, while those ranked below $x^*$ keep the same score.
Hence, for every $z \in X \setminus \{x^*\}$ we have $S_{B}(z,P)=S_{B}(z,P_{X})+ k(n-N_{P_X}(x^*,z))$, therefore, $x^*$ is a cowinner in $P$ if and only if
$S_{B}(x^*,P_{X})+ kn \geq S_{B}(z,P_{X})+ k(n-N_{P_X}(x^*,z))$, which is equivalent to
$k \geq [S_{B}(z,P_X) - S_{B}(x^*,P_X)]/ N_{P_X}(x^*,z)$. (We recall that $N_{P_X}(x^*,z)$ stands for the number of votes in $P_X$
ranking $x^*$ above $z$).
\hfill \qed\\
In words, checking
whether $x^*$ is a possible cowinner boils down to checking, for each
other candidate $z$, whether there are enough votes where $x^*$ is
preferred to $z$ to compensate for the score difference with this
candidate.
This means that possible cowinners with respect to adding any
number of new candidates can be computed in polynomial time \jl{for
Borda}, and more generally for any rule satisfying the
conditions of Lemma \ref{lemma-borda}. Note that
computing possible winners for Borda is
{\sf NP}-hard \cite{XiaConitzer11}, therefore, the restriction of the problem to
candidate addition induces a complexity reduction.
\begin{example}
Take $X = \{a,b,c,d\}$, $n = 4$, and $P_X = \langle bacd, bacd, bacd, dacb\rangle$.
The Borda scores in $P_X$ are $S_B(a,P_X)=8$, $S_B(b,P_X)=9$, $S_B(c,P_X)=3$, and $S_B(d,P_X)=4$, while $N(a,b)=1$, $N(a,c)=4$, $N(a,d)=3$, $N(b,c)=3$, $N(b,d)=3$, $N(c,d)=3$, and for all $x,y$, $N(x,y)=4-N(y,x)$.
Let $\delta(x,z) = S_{B}(z,P_X) - S_{B}(x,P_X)/N_{P_X}(x,z)$.
The following matrix gives the values of $\delta(x,z)$ for the possible pairs of distinct candidates (for the sake of readability, non-positive values are denoted by $\leq 0$).
\[
\begin{array}{|c|cccc|c|}
\hline
\delta(x,z) & a & b & c & d & \max \\
\hline
a & - & 1 & \leq 0 & \leq 0 & 1 \\
b & \leq 0 & - & \leq 0 & \leq 0 & \leq 0 \\
c & + \infty & 5 & - &\leq 0 & + \infty \\
d & 5 & 6 & 1 & - & 6 \\
\hline
\end{array}
\]
\Omit{
Hence the values of $\delta(x,z)$ for each pair $x,z \in X$:
\begin{itemize}
\item $\delta(b,x) = 0$ for each $x \neq b$;
\item $\delta(a,b) = \frac{9-8}{1} = 1$; $\delta(a,c) = \delta(a,d) = 0$;
\item $\delta(c,a) = \frac{8-4}{0} = +\infty$; $\delta(c,b) = \frac{9-4}{1} = 5$; $\delta(c,d) = 0$;
\item $\delta(d,a) = \frac{8-3}{1} =5$; $\delta(d,b) = \frac{9-3}{1} =6$; $\delta(d,c) = \frac{4-3}{1}=1$.
\end{itemize}
}
Applying Proposition \ref{prop-borda}, $b$ is a possible cowinner whatever the value of $k$, $a$ is a possible cowinner
if and only if $k \geq 1$, $d$ is a possible cowinner if and only if $k \geq 6$, $c$ is not a possible cowinner whatever the value of $k$\footnote{This is so because $c$ is always ranked below $a$. We make this intuition clear in Section \ref{related}.}.
Note that for $k \geq 6$, $d$ is a possible cowinner whereas $c$ is not, although $c$ has a higher Borda score than $d$ in $P_X$.
\end{example}
\section{Hardness with a single new candidate}\label{hardone}
Even though we have seen that the possible cowinner problem can be
{\sf NP}-hard for some scoring rules, {\sf NP}-hardness required the addition
of several new candidates. We now show that there exists a scoring
rule for which the possible cowinner problem is {\sf NP}-hard with respect
to the addition of {\em one} new candidate.
The scoring rule we use is very simple: it allows each voter to
approve exactly 3 candidates, and offers 3 different levels of
approval (assigning respectively 3,2,1 points to the three preferred
candidates). Let $r_\Delta$ be the scoring rule defined by the
vector $\vec{s}=\langle 3,2,1,0,\ldots, 0\rangle$ with $m-3$ 0's
completing the vector.
\begin{prop}\label{PropNP-completeVoting}
Deciding if $x^*$ is a possible cowinner for $r_\Delta$ with
respect to the addition of one candidate is {\sf NP}-complete.
\end{prop}
\noindent {\em Proof:~}
This problem is clearly in {\sf NP}. The hardness proof is quite similar
to that of Proposition \ref{PropNP-completeApproval}. Let
$I=(\mathcal{C},A\times B\times C)$ be an instance of \textsc{3-DM}
with $n'\geq 5$ and $\forall z\in A\cup B\cup C$, $d(z)\in\{2,3\}$.
From $I$, we build an instance $\langle N, X, P_X, 1, x^* \rangle$
of the PcWNC problem as follows. The
set $X$ of candidates contains $x^*$, $X_1= \{x'_i,y'_i,z'_i:1\leq
i\leq n'\}$ where $x'_i,y'_i,z'_i$ correspond to elements of $A$,
$B$ and $C$ respectively and a set $X_2$ of dummy candidates. The
set $N$ of voters contains $N_1=\{v^e:e\in\mathcal{C}\}$ and a set
$N_2$ of dummy voters. For each voter $v_i\in N$, we only indicate
the vote for the first three candidates. So, the vote
$V_i=(t_1,t_2,t_3)$ means that candidate $t_i$ receives $4-i$
points. The vote $V_e$ of voter $v^e$ is $(x'_i,y'_j,z'_k)$ where
$e=(a_i,b_j,c_k)\in\mathcal{C}$. The preferences of dummy voters are
such that \jm {$(a)$ the score of the candidates in $X$ satisfies
$\forall x\in X_1$, $S_{\vec{s}}(x,P_X)=3n'+1$,
$S_{\vec{s}}(x^*,P_X)=3n'$ and $\forall x\in X_2$,
$S_{\vec{s}}(x,P_X)\leq 3$ and $(b)$ each voter in $N_2$ ranks at
most one candidate of $\{x'_i,y'_i,z'_i:1\leq i\leq n'\}$ in the
first three positions, and if he ranks one in second position, then
$x^*$ occurs in third position.}
Formally, the instance of the PcWNC problem is built as follows.
The set of voters is $N=N_1\cup N_2$ where
$N_1=\{v^e:e\in\mathcal{C}\}$ and $N_2=N_{A}\cup
N_{B}\cup N_{C}\cup N_{x^*}$, the set of
candidates is $X=X_1\cup X_2\cup \{x^*\}$ where $X_1= X'\cup Y'\cup
Z'$ with $X'=\{x'_1,\dots,x'_{n'}\}$, $Y'=\{y'_1,\dots,y'_{n'}\}$,
$Z'=\{z'_1,\dots,z'_{n'}\}$ and $X_2=X_{A}\cup
X_{B}\cup X_{C}\cup X_{x^*}$. These
sets are defined as follows:\\
\begin{itemize}
\item[$\bullet$] $X_{A}=\{x_i^j:1\leq i\leq n'$,
$1\leq j\leq 2\left(n'-d(a_i)\right)\}$.
\item[$\bullet$] $X_{B}=\{y_i^j:1\leq i\leq n'$,
$1\leq j\leq 2\left(3n'-2d(b_i)+1\right)\}$.
\item[$\bullet$] $X_{C}=\{z_i^j:1\leq i\leq n'$,
$1\leq j\leq 2\left(3n'-d(c_i)+1\right)\}$.
\item[$\bullet$] $X_{x^*}=\{x^*_i:i=1\leq i\leq 2n'\}$.
\end{itemize}
The set of all $X$-votes ${\cal P}_X$ is given by:
\begin{itemize}
\item[$\bullet$] $N_{A}=\{v_{i,j}^{A}:1\leq i\leq n'$,
$0\leq j \leq \left(n'-d(a_i)-2\right)\}\cup
\{v_{i}^{A,j}:1\leq i\leq n'$, $j=1,2\}$. The vote
$V_{i,j}^{A}$ of $v_{i,j}^{A}$ is
$V_{i,j}^{A}=(x'_i,x_i^{2j+1},x_i^{2j+2})$. Note that
$n'-d(a_i)-2\geq 0$. The vote of $v_{i}^{A,j}$ is
$V_{i}^{A,j}=(x_i^{2\left(n'-d(a_i)-1\right)+j},x'_i,x^*)$.
\item[$\bullet$] $N_{B}=\{v_{i,j}^{B}:1\leq i\leq n'$,
$0\leq j \leq 3n'-2d(b_i)\}$. The vote of $v_{i,j}^{B}$
is $V_{i,j}^{B}=(y_i^{2j+1},y_i^{2j+2},y'_i)$.
\item[$\bullet$] $N_{C}=\{v_{i,j}^{C}:1\leq i\leq n'$,
$0\leq j \leq 3n'-d(c_i)\}$. The vote of $v_{i,j}^{C}$
is $V_{i,j}^{C}=(z_i^{2j+1},z_i^{2j+2},z'_i)$.
\item[$\bullet$] $N_{x^*}=\{v_j^{x^*}:0\leq j \leq n'-1\}$. The vote of
$v_j^{x^*}$ is $V_j^{x^*}=(v_{2j+1}^{x^*},v_{2j+2}^{x^*},x^*)$. Note
that $n'-1\geq 0$.
\item[$\bullet$] $N_1=\{v^e:e\in\mathcal{C}\}$. The vote of
$v^e$ is $V_e=(x'_i,y'_j,z'_k)$ where
$e=(a_i,b_j,c_k)\in\mathcal{C}$.
\end{itemize}
We claim that $I$ admits a perfect matching $M\subseteq \mathcal{C}$
if and only if $x^*$ becomes a possible cowinner by adding a new
candidate $y_1$.
Observe that the scores of the candidates in $X$ satisfy:
\begin{itemize}
\item[$(i)$] $\forall x\in X_1$, $S_{\vec{s}}(x,P_X)=3n'+1$.
\item[$(ii)$] $S_{\vec{s}}(x^*,P_X)=3n'$.
\item[$(iii)$] $\forall x\in X_2$, $S_{\vec{s}}(x,P_X)\leq 3$.
\end{itemize}
\jm{Items $(i)$, $(ii)$ and $(iii)$ correspond to the conditions
$(a)$ and $(b)$ described previously.}
For instance, each candidate $x'_i$ from $X_1$ gets respectively 3, 3, and 2 points from the votes $V_e$, $V_{i,j}^A$, and $V^{A,j}_i$, summing up to $3 d(a_i) + 3(n' -d(a_i) -1) +2 = 3n'+1$. The reader can easily check that the conditions also hold for all other candidates.
Let $y_1$ be the new candidate. By construction of this scoring
rule, we must decrease the score of candidates in $X$ which dominate
the score of $x^*$, that is the
candidates of $X_1$ using $(i)$ and $(iii)$.\\
Let $P'$ be a $X\cup \{y_1\}$-profile such that $x^*$ is a cowinner.
Let us focus on candidates in $X'$. In order to reduce
the score of $x'_i$ by 1, we must modify the preference for at least one
voter $v^e$ or $v_{i,j}^{A}$ or $v_{i}^{A,j}$. If we modify it for
some voter in $v_{i}^{A,j}$, then the score of $x'_i$ (with respect
to $v_{i}^{A,j})$ decreases by one if and only if the score of $x^*$ (with
respect to $v_{i}^{A,j})$ also decreases by one. In conclusion, we
must modify the preference of $x'_i$ for at least one voter $v^e$ or
$v_{i,j}^{A}$. By construction, each such voter must put $y_1$ in
top position and then, the score of $y_1$ increases by 3 at each
time. Since there are $n'$ candidates in $X'$, we deduce
$S_{\vec{s'}}(y_1,P')\geq 3n'$; From above remark, we also get
$S_{\vec{s'}}(x^*,P')\leq S_{\vec{s'}}(x^*,P_X)=3n'$. Thus for each
$i\in \{1,\dots,n'\}$, exactly one voter among those of $v^e$ or
$v_{i,j}^{A}$ must put candidate $y_1$ in top position. Finally, if
it is one voter $v_{i,j}^{A}$, then we deduce
$S_{\vec{s'}}(y_1,P')>3n$ because the score of $Y'\cup Z'$ must also
decrease, which is not possible since $y_1$ will then win.
Following a line of reasoning similar to the one developed in the
proof of Proposition \ref{PropNP-completeApproval}, we conclude that
for each $i\in \{1,\dots,n'\}$, exactly one voter among those of
$v^e$ must put candidate $y_1$ in top position. Since the score
of $Y'\cup Z'$ must also decrease by 1, we deduce that $x^*$ is a
possible cowinner if and only if $M=\{e\in \mathcal{C}:$ $y_1$ is in top position
in vote $V_e\}$ is a perfect matching of $\mathcal{C}$ (for the
remaining voters, $y_1$ is put in last position).
\hfill \qed\\
This rule shows that it may be difficult to identify
possible cowinners with a single additional candidate. Giving a
characterization of all rules possessing this property is an open
problem.
\section{Related work}\label{related}
\subsection{The possible winner problem}\label{pwp}
The possible winner problem was introduced in \cite{KonczakLang05}: given an incomplete profile $P = \langle V_1, \ldots, V_n\rangle$ where each $V_i$ is a partial order over the set of candidates $X$, $x$ is a possible winner for $P$ given a voting rule $r$ if there exists a complete extension $P' = \langle V_1', \ldots, V_n'\rangle$ of $P$, where each $V_i'$ is a linear order on $X$ extending $V_i$, such that $r(P') = x$. Possible winners are defined in a similar way for a voting correspondence $C$, in which case we say that $x$ is a possible {\em cowinner} if there exists an extension $P'$ of $P$ such that $x \in C(P')$. Clearly, the possible winner problem defined in this paper is a restriction of the general possible winner problem to the following set of incomplete profiles:
\begin{center}
(Restr) {\em there exists $X' \subseteq X$ such that for every $i$, $V_i$ is a linear order on $X'$}
\end{center}
As an immediate corollary, the complexity of the possible (co)winner problem with respect to candidate addition is at most as difficult as that of the general problem. This raises the question whether (Restr) leads to a complexity reduction for the scoring rules we have considered here.
The possible (co)winner problem for scoring rules has received a significant amount of attention in the last years. Xia and Conitzer \cite{XiaConitzer11} proved that the problem was {\sf NP}-complete for the Borda rule, and more generally for scoring rules whose scoring vector contains four consecutive, equally decreasing values, followed by another strictly decreasing value. Betzler and Dorn \cite{BetzlerDorn09} went further by showing that {\sf NP}-completeness holds more generally for all pure\footnote{A (family of) scoring rules $(r_m)_{m \geq 1}$ is pure if for each $m$, the scoring vector for $m+1$ candidates is obtained from the scoring vector for $m$ candidates by inserting an additional score at an arbitrary position. All interesting families of scoring rules are pure; this is in particular the case for $K$-approval and Borda.} scoring rules, except plurality, veto, and scoring rules whose vector $s^m$ is of the form $s^m = \langle 2, 1,\ldots, 1, 0 \rangle$ for large enough values of $m$. The issue was finally closed by Baumeister and Rothe \cite{BaumeisterRotheECAI10}, who showed that the problem for $s^m = \langle 2, 1,\ldots, 1, 0 \rangle$ is {\sf NP}-complete as well. These results compare to ours in the following way: all our {\sf NP}-hardness results strengthen the known {\sf NP}-hardness results for the general possible winner problem, while our polynomiality results show a complexity reduction induced by (Restr).
Two recent papers give results about the {\sc PcWNC} problem for other voting rules. Xia {\em et al.} \cite{XLM11} give results about the possible (co)winner with new \nicobf{candidates} for other voting rules: they showed that {\sc PWNC} and {\sc PcWNC} are {\sf NP}-complete for Bucklin and maximin, that {\sc PcWNC} is {\sf NP}-complete for Copeland$_0$, and they give several results for approval voting, depending on how the extension a vote is defined. Baumeister {\em et al.} \cite{BaumeisterRR11} generalize our Proposition \ref{PropNP-completeVoting} by showing that the {\sc PcWNC} problem is {\sf NP}-complete for any pure scoring rule of the form $\langle \alpha_1, \alpha_2, 1; 0, \ldots, 0\rangle$; they also give {\sf NP}-completeness results for plurality and 2-approval when voters are weighted.
Results about the {\sc PcWNC} known so far (except our Proposition \ref{PropNP-completeVoting} and its generalization by \cite{BaumeisterRR11}) are summarized in the following table. For the sake of completeness, we also mention the complexity of the other prominent subproblem of the possible cowinner problem, namely unweighted coalitional manipulation.
\begin{center}
{\small
\begin{tabular}{l|l|l|l|}
& general problem & candidate addition & manipulation\\ \hline
plurality and veto & {\sf P} & {\sf P} (Prop. \ref{prop-plura})
& {\sf P} \\
Borda & {\sf NP}-complete\ \cite{XiaConitzer11} & {\sf P} (Prop. \ref{prop-borda})& {\sf NP}-complete\ \cite{BetzlerNiedermeierWoegingerIJCAI11,DaviesEtAlAAAI11} \\
2-approval & {\sf NP}-complete\ \cite{BetzlerDorn09} & {\sf P} \jl{(Coro. \ref{coro-2app})} & {\sf P} \\
$K$-approval, $K \geq 3$ & {\sf NP}-complete\ \cite{BetzlerDorn09} & {\sf NP}-complete\ (Prop. \ref{PropNP-completeApproval}) & {\sf P} \\
Bucklin & {\sf NP}-complete\ \cite{XiaConitzer11} & {\sf NP}-complete\ \cite{XLM11} & {\sf P} \ \cite{XZPCR09}\\
maximin & {\sf NP}-complete\ \cite{XiaConitzer11} & {\sf NP}-complete\ \cite{XLM11} & {\sf NP}-complete\ \cite{XZPCR09}\\
Copeland$_0$ & {\sf NP}-complete\ \cite{XiaConitzer11} & {\sf NP}-complete\ \cite{XLM11} & {\sf NP}-complete\ \cite{FHS08}\\
\end{tabular}
}
\end{center}
Another interesting line of work is the {\em parameterized} complexity of the possible winner problem for scoring rules, which has been investigated in \cite{BetzlerHemmannNiedermeier09}. Among other results, they show that for all scoring rules, the problem is fixed-parameter tractable with respect to the number of candidates (in particular, when the number of candidates is bounded by a constant, the problem becomes polynomial-time solvable). This polynomiality result clearly \nicobf{holds in} the possible winner problem with respect to candidate addition, with some caution: the number of candidates here is the {\em total} number of candidates (the initial ones plus the new ones); this result has practical impact in some situations mentioned in the introduction, such as finding a date for a meeting, where the number of candidates is typically low.
We end this subsection by mentioning other works on the possible winner problem and its variants and subproblems, that are less directly connected to our results. The possible winner problem has also been studied from the probabilistic point of view by Bachrach {\em et al.} \cite{BachrachEtAlAAAI10}, where the aim is to count the number of extensions in which a given candidate is the winner.
Such a probabilistic analysis is highly relevant in candidate-adding situations: given $P_X$, a number $k$ of new candidates, and a prior probability distribution on votes, computing the probability that a given candidate $x \in X$ will be the winner, or that one of the initial (resp. new) candidates will be the winner, is extremely interesting.\footnote{Note that if the voting rule is \nicobf{insensitive to the identity of candidates ({\em i.e.} neutral)}, then although the prior probability that one of the $k$ new candidates will be a cowinner under the impartial culture assumption is at least $\frac{k}{|X|+k}$, this is no longer the case when $P_X$ is known: for instance, let us use plurality and consider the profile $P_X = \langle ab, ab, ab\rangle$, and let the number of new candidates be one. For a third candidate to be a cowinner, he either needs to be placed first in all three votes (which occurs with probability $\frac{1}{27}$), or to be placed first in two votes, but not in the third vote (which occurs with probability $\frac{6}{27}$); therefore the probability that the new candidate is a cowinner in the completed profile is only $\frac{7}{27}$.}
\subsection{Control via adding candidates}\label{control}
The possible winners with respect to the addition of candidates is
highly reminiscent of constructive control by the chair via adding
candidates --- this problem first appeared in
\cite{BartholdiToveyTrick92} and was later studied in more depth
for many voting rules, see {\em e.g.},
\cite{HemaspaandraHR07,FHH09}. However, even if a voting situation
where new candidates are added looks similar to an instance of
constructive control by adding candidates, these
problems differ significantly. In control via adding candidates, the
input consists of a set of candidates $X$, a set of ``spoiler''
candidates $Y$, and a full profile $P_{X \cup Y}$: the chair knows
how the voters would vote on the new candidates; the problem is to
determine whether a given candidate $x^*$ can be made a winner by
adding at most $k \leq |Y|$ candidates from $Y$. In the possible
winner problem with respect to candidate addition, we have to take
into account all possible ways for voters to rank the new
candidates. In spite of their significant differences,
there is a straightforward connection between these problems: if an
instance $\langle N,X, P_{X \cup Y}, x^*, k\rangle$ of control via
adding candidates is positive, then $x^*$ is a possible winner in
$P_X$ with respect to the addition of $k$ new candidates (the voting
rule being the same in both problems).
Bartholdi {\em et al.} \cite{BartholdiToveyTrick92} noted that a voting
rule is immune to control by adding candidates as soon as it
satisfies the Weak Axiom of Revealed Preference (WARP), which
requires that the winner among a set of candidates $W$ to be the winner
among every subset of candidates to which he belongs \cite{Plott76};
formally:
{\em for any $Z \subseteq W$, if $r(P_W) \in Z$ then $r(P_Z) = r(P_W)$.}
This property can be used in a similar way for the possible winner problem with respect to candidate addition. Obviously, if the voting rule $r$ satisfies WARP, then any possible winner from $X$ is a winner for the current profile $P_X$\footnote{In order for the converse to hold, we must add one more condition, such as {\em consensus} (a Pareto-dominated candidate cannot be elected). Then, if the winner for the current profile $P_X$ is $x$, by ranking all new candidates at the bottom of all votes, none of them can be the winner in $P_{X \cup Y}$, and by WARP, no candidate $x' \in X \setminus \{x\}$ can either, therefore $x$ is a possible winner for $P_X$ with respect to candidate addition.}.
Unfortunately, this social-choice theoretic property is very strong: \cite{DuttaJacksonLebreton01} show that a voting rule satisfies this property (there, it is called {\em candidate stability}) and unanimity if and only if it is dictatorial.
\subsection{Cloning}\label{cloning}
Finally, the possible winner problem via candidate addition is closely related to manipulation by candidate cloning. Independence of clones was first studied in \cite{TidemanSCW1987}, further studied in \cite{LLL96,LaslierSCW2000}, and a variant of this property was recently considered from the computational point of view in \cite{ElkindEtAlAAAI2010}. The main difference between $x$ being a possible winner with respect to candidate addition and the existence of a candidate cloning strategy so that $x$ or one of its clones becomes the winner, as in \cite{ElkindEtAlAAAI2010}, is that candidate cloning requires a candidate and its clones to be contiguous in all votes. In other terms, whereas our problem considers the introduction of genuinely new candidates,
cloning merely introduces copies of existing ones.
The complexity of this problem is considered by Elkind {\em et al.} \cite{ElkindEtAlAAAI2010} for several voting rules.
Although the proposed model allows for the possibility of having a bounded number of new clones (via a notion of cost),
most of their results focus on the case of unboundedly many clones. Therefore, to be able to compare their results with ours,
we should first say something about the variant of the possible winner problem with respect to candidate addition,
{\em when the number of new candidates is not known beforehand and can be arbitrarily large}. The definitions of voting
situations and possible winners are straightforward adaptations of Definitions \ref{situation} and \ref{posswin}:
a voting situation is now a triple $\Sigma = \langle N, X, P_X \rangle$ and $x^*$ is a possible cowinner with respect to $\Sigma$
and $r$ if there exists an integer $k$ and a set $Y$ of cardinality $k$ such that there is a $(X \cup Y)$-profile $P$ extending $P_X$
such that $x^* \in r(P)$.
We now give a necessary and sufficient condition for a candidate to be a possible winner, for a class of scoring rules including the Borda rule.
\begin{prop}
\label{pro:unbounded_undom} Let $\mathcal{S}$ be a
collection of scoring vectors $(s^m), m \geq 1$, such that
\begin{itemize}
\item for every $p$, $(s^m_j), 1 \leq j \leq m$ is strictly decreasing;
\item for all $j,j'\in\mathbb{N}$, (1) $\lim_{m\rightarrow\infty}\frac{s_{j}^{m}-s_{j'}^{m}}{s_1}=0$ and
(2) $\lim_{m\rightarrow\infty}\frac{s_{j}^{m}-s_{m-j'}^{m}}{s_1}=1$.
\end{itemize}
Then, $x^*$ is a possible winner w.r.t. $\langle N,X,P_X, + \infty \rangle$
if and only if it is undominated\footnote{We recall that candidate $x$ dominates candidate
$x'$ if every voter ranks $x$ above $x'$, and that a candidate is undominated if no other candidate dominates it.} in $P_X$.
\end{prop}
\noindent {\em Proof:~}
First, suppose $x^*$ is undominated in $P_X$. For any candidate $x_{i}\neq x^*$,
define $\Delta^v_{i}$ as the difference between the score of $x^*$ and
the score of $x_{i}$, divided by $s_1$, in the vote $v$. As in the construction
of Lemma 1, put $k$ new candidates right below $x^*$ in every vote, and let $P'$ be the resulting profile.
As the value of $k$ grows, for any vote $v$ ranking candidate $x^*$ below
$x_{i}$, the value of $\Delta^v_{i}$ will tend towards $0$ (by condition
1). Also, condition 2 ensures that for each vote $v$ ranking $x^*$ above
$x_{i}$, the value of $\Delta^v_{i}$ tends towards $1$.
Because $x^*$ is undominated, such votes always exist for every candidate $x_{i}\neq x^*$.
Therefore, when $k$ grows, $\sum_{v \in P'} \Delta^v_i$ tends towards the number of votes
ranking $x^*$ above $x_{i}$, which is at least 1. This implies that the score of
$x^*$ will be eventually larger than the score of $x_i$, and this is true for every $x_i \neq x^*$,
therefore
$x^*$ will eventually become the winner as $k$ grows. Conversely, suppose $x^*$ is dominated
by some candidate $x_{i}$. Because the scores $(s^m_j), 1 \leq j \leq m$ are strictly decreasing,
the score of $x$ will always remain strictly below the score of $x_i$ in the completion of the profile,
hence $x^*$ is not a possible cowinner.
\hfill \qed\\
Clearly, this large class of voting rules includes Borda, since it satisfies the conditions of Proposition \ref{pro:unbounded_undom}.
However, it does not include plurality, and more generally $K$-approval, which violate condition (1).
Still, a very simple condition can be stated for $K$-approval: a candidate is a possible winner as soon as it is approved at least once.
\begin{prop} When $r$ is $K$-approval, $x^*$ is a possible winner w.r.t. the addition of an unbounded number of new candidates if and only if $S_K(x^*,P_X) \geq 1$.
\end{prop}
\noindent {\em Proof:~}
The condition is obviously necessary. Suppose the condition holds on
a given profile. We extend this profile by taking a set of new
candidates $y_{ij}$ where $1\leq i \leq n$ and $1 \leq j \leq K$.
Consider the $i$-th vote: if $x^*$ is ranked in one the top $k$
positions, put all new candidates at the bottom of the vote.
Otherwise, introduce the new candidates $\{y_{i1}, \dots, y_{iK}\}$
at the top of the vote, and all other new candidates at the bottom.
The score of the new candidates is at most 1, while that of $x_i
\neq x^*$ is at most that of $x^*$ (which is unchanged).
\hfill \qed\\
Note that for $K \geq 2$ this condition does \emph{not} imply that the candidate is undominated (nor vice-versa).
It does obviously when $K=1$, \emph{i.e.}, for plurality. \\
Let us see now how the above results relate to those in \cite{ElkindEtAlAAAI2010}. We first note that in the case of the Borda rule we have the same condition. Indeed one sees intuitively that Lemma \ref{lemma-borda} tells us that for some voting rules (including Borda), introducing new candidates in a contiguous manner, as with cloning, is the best thing to do. For plurality, again the condition is similar in both cases. However, for $K$-approval as soon as $K>1$,
the problem becomes hard in the cloning setting whereas it is easy in our setting with an unbounded number of new candidates.
\section{Conclusion}\label{conclu}
In this paper we have considered voting situations where new candidates may show up during the process. This problem increasingly occurs in our societies, as many votes now take place online (through dedicated platforms, or simply by email exchange) during an extended period of time.
We have identified the computational complexity of computing possible winners for some scoring rules.
Some of them allow polynomial algorithms for the problem (\emph{e.g.} plurality, 2-approval, Borda, veto) regardless of the (fixed) number of new candidates showing up.
For the rules of the $K$-approval family, when $K\geq 3$, the problem remains polynomial only if the number of new candidates is at most $2$.
Finally, we have exhibited a simple rule
where the problem is hard for a single new candidate.
The results address the problem of making some designated candidate a cowinner, which is similar to $x$ being unique
winner under the assumption of the most favourable tie-breaking. In the other extreme case (if we want $x$ to be a strict winner, \emph{i.e.}, to win regardless of the tie-breaking rule), the results are easily adapted: for instance, the inequalities in Proposition \ref{prop-plura} and \ref{prop-borda} become strict. For $K$-approval, the first condition of Proposition \ref{prop-kapp} becomes strict but the second one should now read \begin{small}$S_K(P_X,x) \geq \sum_{x_i \in X}\max(0, S_K(P_X, x_i) - S_K(P_X,x)+1)$\end{small}. As for veto, all other initial candidates need to be vetoed at least once. The hardness proofs can also be readily adapted to the unique winner setting.
A more general treatment would require cumbersome expressions, and is also somewhat problematic since the identities of the new candidates are not known anyway (making it difficult to specify easily a tie-breaking rule on these candidates).
As for future work, a first direction to follow would be to try to obtain more general results for scoring rules, as those obtained by \nicobf{Betzler and Dorn} \cite{BetzlerDorn09} for the general version of the possible winner problem. Extending the study to other families of voting rules, such as rules based on the majority graph, is also worth investigating.
Of course, identifying possible winners is not the end of the story. In practice, as mentioned earlier, one may for instance also be interested in a refinement of this notion: knowing how likely it is that a given candidate will win.
Another interesting issue consists in designing elicitation protocols when the preferences about the `old' candidates are already known. In this case, a trade-off occurs between the storage cost and communication cost, since keeping track of more information is likely to help reduce the burden of elicitation.
\paragraph{Acknowledgements.}We are very much indebted to the reviewers of previous versions of this paper for their extremely detailed and relevant comments.
\bibliographystyle{plain}
|
\section{Introduction}
\label{sec:Intro}
A superfluid-insulator transition in a disordered noninteracting
system of bosons at zero temperature is a special type of quantum
phase transition (QPT) \cite{mac}. Instead of the more conventional
competition between different interactions, it is disorder that causes
a drastic change in the nature of the ground state, thus altering the
physical characteristics of the material. A similar type of transition
from a metal to an insulator, usually called the Anderson localization
transition, was first proposed by Anderson \cite{Anderson}, and has
been extensively studied in electronic systems \cite{lee}. In general,
the approach focus on the conductance behavior as the Fermi level
changes in the vicinity of the \emph{mobility edge\/}, which separates
localized and extended one-particle states. In this context, the lower
critical dimension has been determined to be $d_L=2$, which means that
all the states are localized in one dimension for any finite amount of
disorder. Nevertheless, given that the states in a strictly non
disordered system are extended, there is a clear change of regime when
the disordered strength is reduced to zero, which can be characterized
as a QPT.
In the past decade, enormous progress in the techniques for creating
ultracold atom systems in laboratory settings extended the interest in
the disorder effects and localization to bosonic systems (for recent
reviews, see \cite{fallanir,lagenr,aspectr,palenciar}). For bosons, the
transition is from the \emph{insulator\/} (localized) state to the
superfluid one. It was observed both for laser-speckle disorder
\cite{Billy08} and quasiperiodic optical lattices \cite{Roati08} in
Bose-Einstein condensates of $\,^{87}$Rb and $^{39}$K atoms,
respectively. While speckle disorder comes close to the Anderson-type
\emph{random disorder\/} considered in theoretical approaches,
quasi-periodic lattices present a superposition of the lattice
potential with an incommensurate one and can be viewed as
experimental realizations of pseudo-disorder models like the
Aubry-Andr\'e (AA) model \cite{AA}. The latter also shows superfluid
and localized regimes in one dimension, but the transition between
them occurs at a nonzero critical disorder \cite{AA,ingold}.
Recently, we have investigated numerically the superfluid-insulator
transition in one-dimensional, noninteracting systems of bosons with
these two types of disorder \cite{jardel}. Here we focus on the
scaling properties of the superfluid fraction near the
superfluid-insulator transition, obtaining the relevant critical
exponents. For random disorder, even though the superfluid phase is
destroyed for arbitrarily weak disorder, we show that the transition
can still be described as a quantum critical phenomenon with
well-defined critical exponents and power-law scaling behavior. The
same happens for the AA model, but the universality classes are
different.
Our starting point is a well-known scaling relation for the
singular part of the superfluid density $\rho_s$ close to a quantum
superfluid-insulator phase transition \cite{Fisher89},
\begin{equation}
\label{eq:rhos}
\rho_s \sim |g|^{\nu(d+z-2)},
\end{equation}
where $g$ measures the distance to the quantum critical point (QCP),
$\nu$ is the correlation length exponent (i.e., the correlation length
diverges as $\xi \sim |g|^{-\nu}$ at the QCP), $d$ is the spatial
dimension, and $z$ is the dynamic critical exponent associated with
the QCP. The superfluid density is directly related to the
\textit{helicity modulus\/} \cite{fisherpai}, and can be viewed as a
measure of the system's response to a phase-twisting field. Thus, it
is natural to interpret the correlation length as a phase-coherence
length. In the insulating phase it should coincide with the
localization length, which measures the spatial extent of the wave
functions. This holds also for disordered metals \cite{lee}.
In a finite system, even at criticality, the correlation length is
limited by the system size $L$, and the finite-size-scaling form of
the superfluid density is
\begin{equation}
\label{eq:rhosfinite}
\rho_s \sim L^{-(d+z-2)} F(L/\xi) = L^{-(d+z-2)} F(L |g|^{\nu}) \,.
\end{equation}
The corresponding relation for the superfluid fraction ($f_s=L^d
\rho_s$) is
\begin{equation}
\label{eq:sffscaling}
f_s \sim L^{-(z-2)} F(L |g|^{\nu}) \,.
\end{equation}
This last equation is suitable to determine the critical exponents
$\nu$ and $z$ from a numerical evaluation of $f_{s}$ for various
lattice sizes, as we do in the following.
\section{Anderson-like disorder}
\label{sec:And}
The usual Hamiltonian describing interacting bosons on a lattice is
known as the Bose-Hubbard Hamiltonian, and is given by
\begin{equation} \label{eq:ham}
H = \sum_{i}\varepsilon_i n_i + \Omega\sum_{\langle ij \rangle}(
a^{\dag}_i a_j +a^{\dag}_j a_i ) + \frac{U}{2} \sum_{i} n_i (n_i -
1) \,,
\end{equation}
where $a^{\dag}_i$ and $a^{}_i$ are the creation and annihilation
operators of a boson at the lattice site $i$, $n_i=a^{\dag}_i a^{}_i$
is the corresponding number operator, each site has a single bound
state of energy $\varepsilon_i$, hopping between sites is restricted
to nearest neighbors, with amplitude $\Omega$, and $U$ is a local
repulsive interaction. In the rest of this paper, energies are
measured in units of the tunneling amplitude $\Omega$. An
Anderson-like disorder \cite{Anderson} is introduced by choosing
random local energies with a uniform distribution in the range
$-\Delta/2 \le \varepsilon_i \le \Delta/2$, so that $\Delta$ is a
measure of the disorder strength.
We carried out a thorough numerical analysis of the above model for
the non-interacting case in one spatial dimension. Details of this
numerical study are given in Ref.~\cite{jardel}. We recall one of the
main results reported there, namely that the superfluid fraction for a
lattice of size $L$ obeys the relation $f_s^{} =
\exp(-\Delta/\Delta_L)^{4/3}$, where $\Delta_L$ is a characteristic
disorder strength for suppression of superfluidity, which scales with
the lattice size as $ \Delta_L = C\, L^{-3/2} $. This latter relation
is consistent with the expected value of the critical disorder
strength $\Delta_{c}=0$ for destroying the superfluid phase in the
thermodynamic limit for a one-dimensional system. Furthermore,
defining $g \equiv \Delta - \Delta_{c} = \Delta$, this scaling of
$\Delta_L$ with $L$ is recognized as the finite-size version of the
general relation $\xi \sim |g|^{-\nu}$, immediately yielding the
correlation-length exponent $\nu=2/3$.
Equation~(\ref{eq:sffscaling}) implies that $L^{z-2} f_{s}$ is a
universal function of $L\Delta^{\nu}$. The corresponding plot of our
data for different lattice sizes is shown in Fig.~\ref{fig:sffAnd},
where it is clear that all the data collapse onto a universal
curve. The appearance of $f_s$ alone as the scaling quantity in the
vertical axis means that the dynamic critical exponent is $z=2$. The
scaled variable of the horizontal axis in the collapsed plot confirms
the value $\nu=2/3$ for the correlation-length exponent. Actually,
for the present problem, we are able to determine explicitly the
scaling function $F(x)$ in Eq.~(\ref{eq:sffscaling}). The above
mentioned expression for $f_s^{}$ implies that
$F(x)=\exp(-x^{2}/C^2)$. Since $x=L/\xi=L\Delta^{\nu}$, the value
$z=2$ for the dynamic exponent implies a jump of the superfluid
fraction from 0 to 1 at the transition occurring for $\Delta=0$ in the
limit $L \to \infty$. This jump is reminiscent of that of the helicity
modulus in the two-dimensional $XY$ model \cite{harada}.
\begin{figure}
\begin{center}
\includegraphics[width=0.95\linewidth]{fig1.eps}
\end{center}
\caption{(Color online) Finite size scaling of the superfluid fraction
for Anderson-type disorder. According to Eq.~(\ref{eq:sffscaling}),
the horizontal-axis variable for collapsed curves is $L |g|^{\nu}$,
which gives $\nu=2/3$, while the absence of any rescaling of $f_s$
implies that $z=2$.} \label{fig:sffAnd}
\end{figure}
The value $\nu=2/3$ obtained here for the correlation-length exponent
has not been determined previously, to the best of our knowledge. This
new exponent for the superfluid-insulator transition seems to violate
the inequality $\nu \ge 2/d$, which holds for other disordered systems
\cite{Fisher89}. However, this inequality has been proved only for
interacting systems and for nonzero critical values of the parameter
driving the transition, which is not the case here. On the other hand,
the dynamic exponent $z=2$ implies that the effective dimension of the
quantum phase transition \cite{mac} is $d_{\mathrm{eff}}=d+z=3$. For
disordered interacting bosons, the Bose-glass-to-superfluid transition
is characterized by the relation $z=d$ \cite{Fisher89}. Therefore,
interacting and noninteracting bosons are in different universality
classes with respect to the localization transition. In the
renormalization group language, interaction is a relevant term close
to the disordered non-interacting fixed point.
\section{Aubry-Andr\'e Model}
\label{sec:AA}
\begin{figure}
\includegraphics[width=0.95\linewidth]{fig2.eps}
\caption{(Color online) Finite size scaling of the superfluid fraction
for the AA model. The data collapse in different curves for even
and odd numbers of lattice sites (respectively, lower and upper
curves).} \label{fig:sffAA}
\end{figure}
\begin{figure}
\includegraphics[width=0.95\linewidth,clip]{fig3.eps}
\caption{(Color online) Spectra of the Harper model, highlighting the
spectrum corresponding to the AA model (for a rational approximation
of the golden ratio $\beta=987/610$). Its fractal nature is
illustrated in the bottom by expanding the small box drawn inside
the middle panel. We show the spectrum for $\beta -1$,
which is the same as for $\beta$ according to
Eq.~(\ref{eq:harper}).} \label{fig:btfly}
\end{figure}
The Aubry-Andr\'e model can also be described by the
Hamiltonian~(\ref{eq:ham}), with $U=0$, except that the distribution
of local energies is not random, but periodic with a period
incommensurate with the lattice spacing. These energies are usually
written as
\begin{equation}
\label{eq:harper}
\varepsilon_i = \Delta \cos(2\pi\beta i),
\end{equation}
where $\beta=(1+\sqrt{5})/2$ is the golden ratio and $i$ assumes
integer values from 1 to $L$. This is actually a special case of the
Harper model \cite{harper} for electrons in a two-dimensional lattice
in the presence of a perpendicular magnetic field, for which
Eq.~(\ref{eq:harper}) holds for any value of $\beta$, with different
characteristics of the spectrum for rational or irrational
values. Disorder-like effects here are a consequence of the
incommensurability between the ``external potential'' and the
lattice. Aubry and Andr\'e \cite{AA} proved that for this model
localization occurs only when the strength of the potential $\Delta$
is larger than the critical value $\Delta_c = 2$. For finite lattices,
it is convenient to replace $\beta$ with $\beta_n=F_{n+1}/F_n$, the
ratio of two consecutive Fibonacci numbers, whose limit for $n \to
\infty$ is the golden ratio. Then, the lattice size must be chosen as
$L=F_n$ in order to allow for the use of periodic boundary
conditions. For this kind of finite lattices, the critical value
$\Delta_c = 2$ remains a rigorous result \cite{ingold}, since it
corresponds to a duality between the Hamiltonians in position and
momentum space.
It was shown in Ref.~\cite{jardel} that the superfluid fraction
undergoes a very sharp transition around $\Delta=2$ for essentially
all lattice sizes. This sharpness makes it difficult to directly
extract the correlation-length exponent, as done for random disorder
(Sec.~\ref{sec:And}). Here, we concentrate on a narrow region around
$\Delta_c$, searching for the appropriate scaling variable
proportional to $g \equiv \Delta-\Delta_c = \Delta-2$, and the
appropriate scaling of the superfluid fraction. Our results are shown
in Fig.~\ref{fig:sffAA}. The data collapse onto two universal curves,
for even and odd numbers of lattice sites. Although the scaling
functions are different for these two cases, the critical exponents
for which the curves collapse are the same. In view of
Eq.~(\ref{eq:sffscaling}) we immediately identify the
correlation-length exponent $\nu=1$ from the $x$-axis scaling variable
in Fig.~\ref{fig:sffAA}, and the dynamic exponent $z=2.374\;$ from the
$y$-axis scaling. It was already known \cite{AA} that $\nu=1$ for this
model. Next we discuss the obtained value of $z$ in the light of
properties of the energy spectrum.
The spectrum of the Harper model has been thoroughly studied in the
past \cite{kohmoto1,Sokoloff,kohmoto2,kohmoto3,kohmoto4}. For general
rational values of $\beta$ it is multifractal at $\Delta=2$, yielding
the famous \emph{Hofstadter butterfly\/} \cite{Hofstadter}, shown in
Fig.~\ref{fig:btfly}. There we highlight the case that we are studying
here, with $\beta$ being a rational approximant of the golden ratio.
In particular, the figure shows our result obtained from numerical
diagonalization of the Hamiltonian for a lattice of size $L=610$. The
two bottom panels illustrate the fractal nature of this spectrum.
With our replacement of $\beta$ by a ratio of two Fibonacci numbers,
$\beta_n=F_{n+1}/F_n$, the spectrum is equivalent to the one for
$\bar\beta_n = \beta_n - 1 = F_{n-1}/F_n$, which contains $F_n$ bands
and $F_{n-1}$ gaps. As discussed in detail in
Refs.~\cite{kohmoto2,kohmoto3}, when $F_n = L \to \infty$ the width
$\Delta E_L$ of a given band belonging to the spectrum scales as
$\Delta E_L \sim L^{-\gamma}$, with different regions of the spectrum
associated with different values of $\gamma$ (not to be confused with
the susceptibility critical exponent). In particular, a maximum value
$\gamma_{\mathrm{max}} = 2.374$ corresponds to band-edge states. On
the other hand, the band width is a characteristic energy of the
system and therefore should scale as $\xi^{-z}$, which means that
$\Delta E_L \sim L^{-z}$. Our finding of $z=\gamma_{\mathrm{max}}$ is
in agreement with the relevant state for the zero-temperature
superfluid-insulator transition being the bottom edge of the
lowest-lying band.
\section{Effects of interaction}
\label{sec:ineract}
\begin{figure}
\includegraphics[width=0.95\linewidth]{fig4.eps}
\caption{(Color online) Zero-temperature phase diagram near the
localization transition in the presence of a small interaction. The
color scale indicates values of the superfluid fraction. The
line $U_c(g)$ in the vicinity of the non-interacting QCP is
essentially linear (with a slope close to 0.1) implying that it is
dominated by the analytic part $f(g)$ in
Eq.~(\ref{eq:Uc_gen}).} \label{fig:UxDel}
\end{figure}
The interaction term in Eq.~(\ref{eq:ham}) can be treated as a
relevant field close to the QCP. The knowledge of the dynamic exponent
allows us to generalize the scaling relations close to the
QCP for small but finite $U$. The free energy, for example, is given
by
\begin{equation}
\label{eq:freen}
F_s \propto |g|^{\nu(d+z)} P({U}/{|g|^{\nu z}})\,,
\end{equation}
where again we used the fact that $U$ is an energy, and thus scales
with $\xi^{-z}$. From the above equation, we see that the scaling
contribution to the critical line separating the superfluid and
insulating phases is
\begin{equation}
\label{eq:Uc}
U_c(g) \propto |g|^{\nu z}\,.
\end{equation}
The critical exponents are those associated with the disordered
non-interacting QCP at $U=0$, $\Delta=\Delta_C=2$. Since the product
$\nu z=2.374$ is large we have to take into account analytic
contributions to the shape of the critical line. This line can be
written in general as
\begin{equation}
\label{eq:Uc_gen}
U_c = f(g) + a_{\pm} |g|^{\nu z} \,,
\end{equation}
where $f(g)$ is an analytic function, and $\pm$ refers to the sign of
$g$. Expanding close to the QCP, analytic contributions up to the
second order dominate over the scaling term when $g \rightarrow 0$.
To illustrate this point, in Fig.~\ref{fig:UxDel} we show a phase
diagram close to the noninteracting fixed point at $\Delta_c=2$, for
repulsive ($U>0$) and attractive ($U<0$) interactions. This is a plot
of the superfluid fraction (color scale) as a function of $\Delta$ and
$U$, obtained by diagonalizing the Hamiltonian~(\ref{eq:ham}) for
$N=8$ interacting bosons on a lattice of $L=8$ sites. We can see that
a straight line [i.e., $f(g) \sim g$] is a very good approximation to
the boundary between the superfluid and localized regions. Even though
the transition is smoothed out by the small lattice size, it is worth
mentioning that the value of $f_s$ at the critical point
$(U=0,\Delta=2)$ is compatible with the lower curve of
Fig.~\ref{fig:sffAA} for $L=8$.
\section{Conclusions}
\label{sec:concl}
We studied the superfluid-insulator transition for bosons on a
one-dimensional lattice, both with random disorder and the pseudo
disorder described by the Aubry-Andr\'e model, the two prototype
models employed in the investigation of localization for ultracold
atoms in optical lattices. Using a finite-size-scaling analysis of the
superfluid fraction, we obtained the critical exponents characterizing
this transition. The superfluid fraction yields the
correlation-length exponent $\nu$ and the dynamic critical exponent
$z$. For random disorder we found $\nu=2/3$ and $z=2$, while for the
AA model the results are $\nu=1$ and $z=2.374$. The other critical
exponents can be obtained from the quantum hyperscaling relations
\cite{mac} $2-\alpha=\nu(d+z)$ and $2 \beta=\nu(d+z-2+\eta)$. These
two models fall into different universality classes, which is not
surprising since the critical disorder strength for the
superfluid-insulator transition is zero for Anderson-like disorder and
nonzero for the AA model, which also exhibits a multifractal energy
spectrum at the QCP.
It is interesting to observe that the scaling form of the free energy
for nonzero temperature $T$ can be used to determine the
thermodynamic behavior close to the superfluid-insulating QCP. From
it, a general dependence of the specific heat with temperature is
obtained \cite{Fisher89}, with the form $C \sim T^{d/z}$.
Our brief discussion of interaction effects in the AA model shows that
the critical point moves to stronger disorder for repulsive
interaction, and to weaker disorder in the attractive case. Regions of
the phase diagram that correspond to localized and superfluid regimes
in the thermodynamic limit are separated by a line that is
approximately linear, reflecting the dominance of nonsingular
contributions.
\subsection*{Acknowledgments}
This work was supported in part by CNPq, Conselho Nacional de
Desenvolvimento Cient\'ifico e Tecnol\'ogico (Brazil). M.A.C. also
thanks FAPERJ, Funda\c{c}\~ao de Amparo \`a Pesquisa do Estado do Rio
de Janeiro, for partial financial support.
|
\section{Introduction}
The properties of magnetic turbulence in interarm versus arm regions
are distinctly different, as shown by Faraday rotation measure (RM)
structure functions determined separately for the two regimes. The
interstellar magnetic field (ISMF) is coherent over 100 pc scales in
interarm regions, and random over spatial scales of less than 10 pc in
arm regions \cite{Haverkorn:2006}. Which of these scenarios dominates
for the ISMF in the solar vicinity? The heliosphere, and its ribbon of
energetic neutral atoms (ENAs) discovered by the Interstellar
Boundary Explorer (IBEX) spacecraft \cite{McComas:2009sci}, probe the
direction and strength of the very local ISMF in a single spatial
location. Over intermediate scales, the position angles of starlight
polarized by magnetically aligned interstellar dust grains trace the
ISMF direction. Pulsar RM data give the field direction for spatial
scales of 100--300 AU. As discussed in Frisch et
al. \cite{Frisch:2011ismf2}, together these data and cosmic ray
anisotropies suggest the ISMF near the Sun is similar to expectations
for interarm regions.
\section{IBEX Ribbon, Heliosphere, and Local ISMF}\label{sec:ibex}
IBEX measures ENAs formed by charge-exchange (CEX) between
heliospheric ions and neutral gas from the surrounding ISM, which
flows through and around the heliosphere at a velocity of $\sim 26$
\hbox{km s$^{-1}$}\ from the direction $\ell$,$b$$\sim 4^\circ, 15^\circ$. Among the sources of the ions creating the 0.01--6 keV ENAs
detected by IBEX are the post-shock solar wind, pickup ions, and
non-thermal energetic ions in the inner heliosheath, secondary ions
created by CEX between interstellar protons and ENAs escaping beyond
the heliopause, and possibly pickup ions inside of the solar wind
termination shock. The first ENA full skymap, collected early 2009,
revealed an unexpected "ribbon" of ENAs that traces heliosphere
asymmetries created by the angle between the ISMF and velocity of
inflowing neutral interstellar gas
\cite[Fig. \ref{fig:1},][]{McComas:2009sci,Funsten:2009sci,Schwadron:2009sci}.
The ribbon is visible in the energy range $\sim 0.2 - 6 $ keV and
forms a nearly complete circle in the sky, that is offset by $\sim
75^\circ$ from the ribbon arc center
\cite{Schwadronetal:2011sep}. Comparisons between heliosphere
models (see below) and the ENA ribbon revealed that the ribbon is seen
towards sightlines that are perpendicular to the ISMF draping over the
heliosphere \cite{Schwadron:2009sci}. Seven possible formation
mechanisms for the ribbon have been discussed
\cite{McComasetal:2010var,Schwadronetal:2011sep,McComas:2009sci,GrzedzielskiBzowski:2010ribbon}.
In order to link the ribbon directly to the ISMF draping over the
heliosphere, the mechanism based on CEX between interstellar H$^\circ$
and energetic ions beyond the heliopause is assumed to be correct
\citep{Heerikhuisen:2010ribbon,Chalov:2010ribbon}. The ribbon
location is stable over the first several skymaps
\cite{McComasetal:2010var}. The ISMF direction at the heliosphere is
then given by the center of the ribbon arc, located at $\ell$=33\hbox{$ ^\circ$},
$b$=55\hbox{$ ^\circ$}\ (see Table 1 for ecliptic coordinates).
\input{frischfig1}
The best match between the predicted and observed ribbon locations occurs for
ENAs produced roughly 100 AU beyond the heliopause
\cite{HeerikhuisenPogorelov:2011}.
The surrounding cloud is $\sim 22$\% ionized \cite[$n(\mathrm{p}) \sim 0.07$ \hbox{cm$^{\rm -3}$},][]{SlavinFrisch:2008}.
The ribbon formation mechanism requires that ENAs escaping the heliosphere
CEX with interstellar protons piled up against the heliopause,
creating a secondary pickup ion population in the outer heliosheath.
The secondary pickup ion must CEX quickly with an interstellar H$^\circ$
atom (before magnetic turbulence disrupts the ring-beam distribution),
to create an ENA detected preferentially in directions perpendicular
to the ISMF. The relative roles of large- and small-scale magnetic
turbulence in the outer heliosheath are not presently understood, so
this mechanism is still under study
\cite{Florinski:2010ribbon,GamayunovZhang:2010ribbon}. A recent
model
\cite{HeerikhuisenPogorelov:2011} reproduces the ribbon for
an ISMF strength of $\le 3$ \hbox{$ \mu{\rm G}$}, with the pole directed from
$\ell$=33\hbox{$ ^\circ$}, $b$=54\hbox{$ ^\circ$}\ (or $\lambda,\beta =
222^\circ,39^\circ$). A local ISMF strength of $\sim 3$ \hbox{$ \mu{\rm G}$}\ is
found from photoionizaton models of the local interstellar cloud (LIC)
around the heliosphere, and the assumption of pressure equilibrium
between the ISMF and thermal gas
\cite{SlavinFrisch:2008}.
MHD heliosphere models are constructed to reproduce the
heliospere asymmetries induced by the $\sim 45^\circ$ angle between
the ISMF direction and gas flow direction
\citep[][]{Pogorelovetal:2009asymmetries,Prestedetal:2010,Opheretal:2007,Ratkiewiczetal:2008}.
The ISMF direction in these models is generally directed downwards
through the ecliptic plane in order to reproduce asymmetries such as
the $\sim 5^\circ$ offset between the central directions of the flows
of interstellar H and He into the heliosphere, and the difference
between the solar wind termination shock distances encountered
by Voyager 1 (94 AU) versus Voyager 2 (84 AU).
If the IBEX ribbon is formed beyond the heliopause through CEX with
secondary pickup ions, then the ribbon is an extraordinary diagnostic
of the local ISMF and neutral densities. Models show that variations
in interstellar properties of $\sim 15$\% lead to pronounced
differences in the location and width of the ribbon because of the
interplay between magnetic, plasma and neutral pressures
\cite{Frisch:2010next,HeerikhuisenPogorelov:2011}.
Two aspects of the ribbon data are notable: (1) The imprint of the
solar wind on ENA fluxes and the ribbon is apparent in flux decreases
of $\sim 15$\% between the first and second skymaps, spaced by six
months \cite{McComasetal:2010var}. The travel time for a 1 keV solar
wind proton to pass the Earth, the solar wind termination shock, and
then return to IBEX as an ENA, is over two years. This time lag
suggests that the decrease in 0.7--1.1 keV fluxes in the second map is
from the unusually low solar wind pressure of the recent solar cycle
minimum. (2) The energy distribution of global ENAs shows an overall
power law spectra, $E^{-\kappa}$, with \hbox{$\kappa$}\ ranging between 1.28 and
2.04 for different regions of the sky and spectral intervals
\cite{Funsten:2009sci}. The power-law distribution of ribbon ENAs becomes softer above a
"knee" in the spectrum between 1--4 keV, which occurs at higher
energies for higher latitudes
\cite{Schwadronetal:2011sep}.
\section{Local ISMF direction from polarized starlight}
The galactic ISMF was first mapped in the 1940's using polarized
starlight, which showed patterns of linear polarization related to
dust extinction. In the diffuse ISM, dust grain alignment occurs when
magnetic torques align asymmetric rotating charged dust grains,
producing a birefringent interstellar medium with lowest opacities parallel to the
ISMF direction. This gives linear polarization that is parallel to,
and traces, the ISMF direction \cite{Roberge:2004}. For sightlines
perpendicular to the ISMF direction in a uniform medium, polarization
will increase with distance. Polarizations do not provide the
strength or polarity of the ISMF, and polarization is reduced by
foreshortening towards the poles of the field.
\input{frischfig2}
During the early 1970's, Tinbergen
\cite{Tinbergen:1982} conducted a sensitive (for that time) study of
the interstellar polarizations of stars within $\sim 40$ pc of the
Sun, and found a patch of dust in the south and close to the Sun.
Based on interstellar absorption lines towards $\alpha$ Cen at 1.3 pc,
and polarization towards 36 Oph, the magnetically aligned dust must be
in the "G-cloud" within 1.3 pc \cite{Frisch:2011araa}. Recent
polarization data confirm that local interstellar polarizations are
significantly weaker in the northern hemisphere than the southern
hemisphere.
We have analyzed available interstellar polarization data, including
six recent sets of sensitive data, to determine the best fitting ISMF
direction for a 180\hbox{$ ^\circ$}\ diameter region that is centered on the
heliosphere nose and coinciding roughly with the G-cloud
(\cite{Frisch:2010ismf,Frisch:2011ismf2}. Although we utilize
polarization towards stars out to 40 pc, most of the polarization
should be formed in the region with the highest concentrations of gas
within $\sim 15$ pc \cite{Frisch:2011araa,Frisch:2011ismf2}. These
data indicate a local ISMF direction with a pole towards
$\ell$,$b$$\sim 38^\circ, 23^\circ$, with uncertainties of $\sim \pm
35^\circ$ (Fig. \ref{fig:2}, Table 1,
\cite{Frisch:2010ismf,Frisch:2011ismf2}). This direction is centered
approximately 35\hbox{$ ^\circ$}\ from the center of the IBEX ribbon arc. The
small difference between the two directions may indicate turbulence in
the local ISMF, for example between the ISMF directions of the G-cloud
(within 1.3 pc) and the LIC in which the Sun is embedded. In several
regions, the flow of interstellar gas past the Sun shows merging or colliding
clouds \cite{RLIV:2008vel}, such as the two clouds in front of Sirius
\cite{FrischMueller:2011ssr}, which could drive local magnetic
turbulence.
\input{frischtable}
\section{The Loop I superbubble and local ISMF}
Loop I is an evolved but reheated superbubble that dominates the
northern hemisphere because of it's large angular radius. It appears as
a distinct imprint on the ISMF within 100 pc that is detected in polarized
starlight, polarized radio continuum, Faraday rotation, and Zeeman
splitting in the HI filaments defining the shell circumference
\cite{Frisch:1995rev}. All existing spherically symmetric models of the radio
continuum Loop I place the Sun in the superbubble rim. A recent study
of the intensity of 1.4 GHz emission in this region derived two
subshells of Loop I containing enhanced synchrotron emissivity from
swept up magnetic fields, and placing the Sun in the rim of the S1
subshell \cite{Wolleben:2007,Frisch:2010s1}. The direction of the
bulk flow of the "cluster of local interstellar clouds" (CLIC) past
the Sun has an upwind direction of
$\ell$,$b$=335\hbox{$ ^\circ$},--7\hbox{$ ^\circ$}\ (\cite{FrischMueller:2011ssr}, after correcting for the solar
velocity through the local standard of rest, LSR),
near the center of the S1 shell,
If the CLIC is part of the nearside of an evolved superbubble shell
associated with Loop I \cite{Frisch:1995rev,Frisch:2011araa}, then
both the polarity and direction of the local ISMF could be ordered
over the $\sim 78$ pc radius of the S1 subshell. The angle between
the local ISMF direction and the LSR CLIC velocity is
68\hbox{$ ^\circ$}--78\hbox{$ ^\circ$}, depending on whether the comparison is with the
polarization axis or ribbon arc center. This large angle suggests
that the very local ISMF is roughly parallel to the local surface of
the expanding Loop I shell.
\section{ISMF from pulsars in the Local Bubble }
The ratio of pulsar Faraday rotation measures (RM) and dispersion
measures provides an electron-density weighted measurement of the ISMF
component parallel to the sightline, and the field polarity. Salvati
(2010) performed a best-fit to the rotation and dispersion measures of
four pulsars within 160--300 pc, and in the very low density third
galactic quadrant corresponding to the interior of the Local Bubble.
He found a best-fit direction of
$\ell$$\sim5^\circ$, $b$$\sim 42^\circ$, and a field strength of
$|B|$=3.3 \hbox{$ \mu{\rm G}$}\ (Table 1). This direction is within 23\hbox{$ ^\circ$}\ of
the center of the IBEX ribbon arc, and 33\hbox{$ ^\circ$}\ of the best-fitting
ISMF direction. The polarity of this ISMF is directed towards the
pole of $\lambda$,$\beta$=232\hbox{$ ^\circ$},18\hbox{$ ^\circ$}, indicating that it is directed
upwards through the plane of the ecliptic.\footnote{Note that the
rotation measure is positive when the ISMF is directed towards the
observer, by definition.} Wolleben \cite{Wolleben:2007} finds a
similar RM polarity in this direction, although variations over small
angular scales are seen.
\section{Galactic cosmic rays and the local ISMF}
An unusual indicator of the ISMF affecting the heliosphere is provided
by GeV--TeV galactic cosmic rays (GCR). Cosmic rays at sub-TeV
energies have gyroradii less than several hundred AU in a 1 \hbox{$ \mu{\rm G}$}\
magnetic field, which is comparable to heliosphere dimensions. GCRs
with energies $\le 10^3$ GeV show pronounced spatial asymmetries
\cite{Nagashimaetal:1998,Halletal:1999gcr,LazarianDesiati:2010},
including a sidereal distribution with a southern GCR deficit, and a
broad excess emission region attributed to the heliotail. The broad
tail-in excess at 500--1000 GeV is centered at ecliptic coordinates of
$\lambda$,$\beta$\ of 71\hbox{$ ^\circ$},--42\hbox{$ ^\circ$}\ to 90\hbox{$ ^\circ$},--53\hbox{$ ^\circ$}, and is skewed
with respect to ecliptic latitude and covers the direction opposite to
the ISMF defined by the ribbon arc ($\lambda$,$\beta$=41\hbox{$ ^\circ$},--39\hbox{$ ^\circ$},
Table 1). This tail-in excess also coincides with the axis of the
ISMF determined from the polarization data,
$\lambda$,$\beta$=83\hbox{$ ^\circ$},--37\hbox{$ ^\circ$}. The minimum in tail ENA fluxes
is 44\hbox{$ ^\circ$}\ west of the downwind gas-flow direction
\cite{Schwadronetal:2011sep}, and it overlaps the excess of sub-TeV
particles. Evidently the GCR excess could form either in deep tail
regions or indicate GCRs streaming along the local ISMF or S1 shell.
Lazarian and Desiati (2010) have modeled the tail-in sub-TeV excess
\nocite{LazarianDesiati:2010}
as due to the stochastic acceleration of particles in the
time-variable magnetic field of the heliotail.
\section{Summary}
The summary of data in the previous sections shows that the direction
of the ISMF appears to be relatively constant over scales of five
orders of magnitude, with warping (large-scale turbulence) or
variations in the field direction typically less than
30\hbox{$ ^\circ$}--40\hbox{$ ^\circ$}\ (Table 1). The agreement between ISMF direction
from pulsars in the Local Bubble, starlight polarizations, the IBEX
ribbon, and heliosphere models, is remarkable considering that the
spatial scales of these estimates differ by five orders of magnitude.
The puzzle is that opposite polarities are inferred for the ISMF
shaping the heliosphere and the pulsar RM data, downwards through
the ecliptic plane in the first case and upwards through the ecliptic
plane in the second case. It could be pure coincidence that the Local
Bubble ISMF direction from the pulsar data and from heliosphere
measurements and models are similar. However the galactic magnetic
field is known to be ordered over kiloparsec spatial scales in low
density interarm regions of the Galaxy such as around the Sun
\cite{JinLinHan:2009}. Perhaps it is not so surprising to see
uniformity of the magnetic field direction over local scales of 100
pc. If the ISMF traced by the IBEX ribbon, polarization data, Loop I
shell, and pulsar data are related, it would mean that Loop I is a
highly asymmetric evolved superbubble, with the Sun in the segment
most distant from the superbubble source and most closely linked to
the interarm field direction.
\begin{theacknowledgments}
The author thanks Paolo Desiati for pointing out the usefulness of
cosmic ray TeV asymmetries for understanding the relation between the
heliosphere and local ISMF. This work was supported through grant
NNX09AH50G to the University of Chicago, and by the Interstellar
Boundary Explorer Mission as a part of NASA's Explorer Program.
\end{theacknowledgments}
|
\section{Introduction}
The quasiparticles in graphene\cite{Geim} obey a linear dispersion
law $\omega= v_F k$ ($v_F\approx c/300$ is a Fermi velocity, $c$ is
a speed of light) at energies less than $2$ eV. Graphene's $2+1$ -
dimensionality and quasi-relativistic Dirac model for its
quasiparticles make it possible to derive the Casimir effect
properties of graphene systems from general constraints and
principles of quantum field theory.
Casimir effect in graphene systems was studied in different papers
\cite{Dobson}-\cite{Ali}. Finite-temperature results were obtained
in Refs. \cite{Gomez} and \cite{Mar1}. In the current paper we
follow the formalism developed in Ref.\cite{Mar1}, an alternative
derivation of the reflection coefficients is presented, the relation
to Feynman diagrams is discussed. First we derive the expressions
for the components of the polarization operator of quasiparticles in
a graphene layer at finite temperature and derive the reflection
coefficients of a flat graphene layer from the solutions of the
boundary problems for vector potentials. The free energy is given
then in two equivalent forms: in terms of reflection coefficients
and closed Feynman diagrams. Finally we study the exceptional
properties of the free energy of a flat graphene layer -- parallel
flat metal system at finite temperatures.
We use the coordinates $x^3$ and $z$ interchangeably throughout the
paper. When needed we select the coordinate $y$ along a direction of
the wavevector ${\mathbf q}=(q^1,q^2)$ (longitudinal direction) and the
coordinate $x$ along a transverse direction.
We use
$\hbar=c=k_B=1$.
\section{Action and polarization operator}
The model is described by the following classical action (assuming
graphene plane lying at $x^3=0$)
\begin{equation}
S= -\frac14 \int d^4x\, F^2_{\mu\nu}+
\int d^3x
\bar\psi \slashed{D} \psi
\label{action}
\end{equation} with
\begin{equation*}
\slashed{D} = (i\partial_0-\mu-eA_0)\gamma_0
+v_F[\gamma^1\(i\partial_1-e{A}_1\) +\gamma^2\(
i\partial_2-eA_2\)] -m\,.
\end{equation*}
Here $\mu$ is a
chemical potential, $m$ is a mass gap of quasiparticle excitations,
$v_F\approx 1/300$ is a Fermi velocity. Since there are $N=4$
species of fermions in graphene, the gamma matrices are in fact
$8\times 8$, being a direct sum of four $2\times 2$ representations
(with two copies of each of the two inequivalent ones),
$\gamma_0^2=-(\gamma^{1,2})^2=1$. The Maxwell action is normalized
in such a way that
\begin{equation}
e^2\equiv 4\pi\alpha =\frac{4\pi}{137}.
\label{e}
\end{equation}
In Minkowski space the one-loop polarization operator can be
expressed in momentum space as
\begin{equation} \Pi^{mn}(p_0,{\bf p})
=ie^2 \int\frac{dq_0d^2 {\bf q}}{(2\pi)^3}\,\,
{\rm tr}\( \hat S(q_0,{\bf q})\tilde\gamma^m \hat S(q_0-p_0,{\bf q}-{\bf p})\tilde\gamma^n\),
\label{Pi_expl}
\end{equation}
where the propagator of the quasiparticles in graphene reads
\begin{equation}
\hat S (q_0,{\bf q}) \equiv \slashed{D}^{-1}\vert_{A_\mu=0} =
-\frac{(q_0+\mu)\gamma_0-v_F\slashed{q}-m}{(q_0+\mu+i\epsilon {\rm\ sgn}q_0)^2-v_F^2{\bf q}^2-m^2}.
\label{hat S}
\end{equation}
Note that for $\mu=m=0$ the pole of the propagator yields the linear
dispersion law for quasiparticles in graphene: $q_0 = v_F |{\bf
q}|$. Here ${\mathbf q}=(q^1,q^2)$, $\slashed{q}=\gamma^1q_1+\gamma^2q_2$.
Vectors with tilde are rescaled by multiplying the spatial
components with $v_F$, i.e., $ \tilde p^j\equiv\eta^j_i p^i=(p_0,v_F
{\mathbf p})$, $\eta=\textrm{diag}(1,v_F,v_F)$.
To introduce the temperature in (\ref{Pi_expl}) we perform the
rotation to the Matsubara frequencies
\begin{equation}
i \int dq_0 \rightarrow - 2 \pi T \sum_{k=-\infty}^{\infty}, \qquad
q_0 \rightarrow 2\pi i T(k+1/2) ,
\label{Mats_prescr}
\end{equation} use the Feynman parametrization
$$
\frac1{ab}=\int_0^1 \frac{dx}{(xa+(1-x)b)^2}
$$
and subsequently change the variables in (\ref{Pi_expl}) in the
spatial part of the loop--integration: ${\bf q}\to {\bf q}+x{\bf
p}$. Then we come to
\begin{multline}
\Pi^{00}=-2 e^2 T N
\sum_{k=-\infty}^{\infty}\int_0^1dx \int\frac{d^2
{\mathbf q}}{(2\pi)^2}\frac{M_0^2+(q_{0k}+\mu)(q_{0k}+\mu - p_0)}
{\[(q_{0k}+\mu-xp_0)^2-\Theta^2\]^2} = \\
- \frac{4\alpha T N}{v_F^2} \sum_{k=-\infty}^{\infty}\int_0^1dx
\int_{\Theta_0}^{+\infty} d\Theta \, \Theta \Biggl(
\frac{\Bigl(\Theta^2 + p_0^2 x(1-x) -2 v_F^2 {\mathbf p}^2 x(1-x)
\Bigr)}{\[(q_{0k}+\mu-xp_0)^2-\Theta^2\]^2 } + \\
\frac{(q_{0k}+\mu)(q_{0k} + \mu -
p_0)}{\[(q_{0k}+\mu-xp_0)^2-\Theta^2\]^2}\Biggr) \, , \label{Pol00}
\end{multline} where $M_0^2=m^2+ v_F^2{\mathbf q}^2-x(1-x) v_F^2{\mathbf p}^2$, and
\begin{align}
\Theta^2 &=m^2+ v_F^2 {\mathbf q}^2-x(1-x)(p_0^2-v_F^2 {\mathbf p}^2) , \\
\Theta_0 &= \sqrt{m^2-x(1-x)(p_0^2-v_F^2{\mathbf p}^2)} . \label{th0}
\end{align}
In equation (\ref{Pol00}) we introduced the integration variable
$\Theta$.
In analogy we get
\begin{equation}
\Pi_1^1 + \Pi_2^2 = - 4\alpha T N
\sum_{k=-\infty}^{\infty}\int_0^1dx \int_{\Theta_0}^{+\infty}
d\Theta \, \Theta \, \frac{2m^2 - 2(q_{0k}+\mu)(q_{0k}+\mu -
p_0)}{\[(q_{0k}+\mu-xp_0)^2-\Theta^2\]^2}
\end{equation}
Summation over the fermion Matsubara frequencies can be made
explicitly by making use of the identities \begin{multline}
\sum_{k=-\infty}^{\infty} \frac1{\[(2 \pi i
T(k+1/2)-b)^2-\Theta^2\]^2}= \\
-\frac{1}{16 \Theta^3 T^2}
\(
\Theta {\rm\ sech}^2\(\frac{\Theta+b}{2T}\)
-2 T \tanh\(\frac{\Theta+b}{2T}\)
\) +(\Theta\to-\Theta) ,
\end{multline}
\begin{multline}
\sum_{k=-\infty}^{+\infty} \frac{(2 \pi i T(k+1/2) +\mu)(2 \pi i
T(k+1/2) +\mu - p_0) }{\[(2 \pi i T(k+1/2)-b)^2-\Theta^2\]^2} = \\
-\frac{1}{16 \Theta^3 T^2} \Biggl( {\rm\ sech}^2\(\frac{\Theta+b}{2T}\)
\Bigl(\Theta^3 + \Theta^2 (2x-1)p_0 - \Theta p_0^2 x(1-x) \Bigr) + \\
+ 2 T \tanh\(\frac{\Theta+b}{2T}\)\Bigl( \Theta^2 + p_0^2
x(1-x)\Bigr)
\Biggr) +(\Theta\to-\Theta), \label{Id2}
\end{multline}
where in (\ref{Id2}) we substituted $b =p_0 x- \mu$.
To perform the integration over $\Theta$ it is convenient to use the
identity $\partial_s \tanh s={\rm sech}^2 s $. Finally we arrive at
the following representation for $\Pi_{00}$ and $ \Pi_{{\rm
tr}}\equiv\Pi_m^m$:
\begin{equation}
\Pi_{{\rm tr}, 00}
=-\frac{2 N \alpha T}{v_F^2} \int_0^1dx \(
f_{tr,00}\tanh\frac{\Theta_0+b}{2 T}
-
\ln\(2\cosh\frac{\Theta_0+b}{2 T}\)
+(\Theta_0\to-\Theta_0) \)
\label{Pi_00} \end{equation} where $\Theta_0\equiv
\sqrt{m^2-x(1-x)(p_0^2-v_F^2{\bf p}^2)}$, $b =p_0 x- \mu$, and
\begin{eqnarray}
f_{00}&=&\frac{-2 v_F^2{\bf p}^2x(1-x) - p_0(1-2x)\Theta_0+
2\Theta_0^2}{4 T \Theta_0},
\label{f_00}\\
f_{\rm{tr}}&=&\frac{2m^2 v_F^2 + 2 x(1-x) v_F^2 p_3^2}{4 T
\Theta_0}- \nonumber \\
& &-\frac{p_0(1-2v_F^2) (1-2x)- 2(1-v_F^2) \Theta_0}{4T}.
\label{f_tr}
\end{eqnarray}
Here $p_3^2\equiv p_0^2-{\mathbf p}^2$. We remind that $N$ is the number of
fermion species, $N=4$ for graphene. Parity-odd contributions to the
polarization tensor cancel out between different species, while the
parity-even contributions add up.
\section{Reflection coefficients}
In this section we find reflection coefficients for transverse
electric (TE) and transverse magnetic (TM) modes. The equations
\begin{equation}
\partial_\mu F^{\mu\nu} +\delta(z) \Pi^{\nu\rho}A_\rho =0 \quad
\end{equation}
lead to the conditions
\begin{equation}
\partial_z A_m |_{z=+0} - \partial_z A_m |_{z=-0} = \Pi_{mn} A^n|_{z=0} ,
\label{SQ}
\end{equation}
where \cite{Zeitlin} \begin{equation}
\Pi^{mn}=\frac1{v_F^2}\eta^m_j \Bigl(
\Pi^{ji}_0 A(p_0,{\bf p})
+p_0^2 \Pi^{ji}_u B(p_0,{\bf p})
\Bigr) \eta_i^n
\label{Pi gen tilda}
\end{equation}
\begin{equation}
\Pi^{ji}_0
=g^{ji}-\frac{\tilde p^j\tilde p^i}{\tilde p^2},\quad
\Pi^{ji}_u
=\frac{\tilde p^j\tilde p^i}{\tilde p^2}-\frac{\tilde p^j u^i + u^j \tilde p^i}{(\tilde pu)}
+\frac{u^ju^i}{(\tilde pu)^2}\tilde p^2 ,
\end{equation}
$u^j=\delta^{j 0} $ and $i,j= 0, 1, 2$. Here $A$, $B$ are scalar
functions.
Let's consider the condition
\begin{equation}
\partial_0 A^0 + \partial_x A^x + \partial_y A^y = 0. \label{gauge}
\end{equation}
In fact, the condition (\ref{gauge}) is quite convenient for a
description of transverse electric and transverse magnetic modes of
the propagating electromagnetic wave.
A nonzero $A_x$, the condition $\partial_x A_x=0$ and the conditions
$A_y=A_z=A_0=0$ describe the propagation of a TE electromagnetic
wave (the electric field is parallel to the surface $z=0$) since
$E_x \sim A_x$. Here the direction perpendicular to the wave vector
of the electromagnetic wave under consideration $(0, p_y, p_z)$ is
denoted by $x$.
For the TE wave we have:
\begin{align}
A_x &= e^{i p_y y} e^{i p_z z} + r_{TE} e^{i p_y y}e^{-i p_z z}
\quad \text{for} \quad z<0 \label{TE11}
\\ A_x &= e^{i p_z z} e^{i p_y y} t_{TE} \quad \text{for} \quad z>0 \label{TE2}
\end{align}
and
\begin{equation}
\Pi_{xn} A^n = A_x A(p). \label{T1}
\end{equation}
Here $p_z^2=p_0^2-p_y^2$. From the continuity of potentials at $z=0$
we obtain $1+r_{TE}=t_{TE}$. Now one substitutes (\ref{TE11}) and
(\ref{TE2}) into (\ref{SQ}) and uses (\ref{T1}) to obtain:
\begin{equation}
r_{TE} = \frac{A(p)}{2ip_z - A(p)} \label{rte2}
\end{equation}
The conditions $A_x=A_z=0$, $p_0 A_0=p_y A_y$ describe the
transverse magnetic (TM) wave. This choice of vector potentials
describes TM wave since $E_z \sim
\partial_z A_0 $ or $B_x \sim \partial_z A_y$.
For $A_0$ we have:
\begin{align}
A_0 &= e^{i p_y y} e^{i p_z z} + r_{A_0} e^{i p_y y}e^{-i p_z z}
\quad \text{for} \quad z<0 \label{TM1}
\\ A_0 &= e^{i p_z z}e^{i p_y y} t_{A_0} \quad \text{for} \quad z>0 \label{TM2}.
\end{align}
and
\begin{equation}
\Pi_{0n} A^n = \bigl(A(p) - B(p) v_F^2 p_y^2 \bigr) p_z^2/\tilde
p_z^2 A_0 ,\label{T3}
\end{equation}
where $\tilde p_z^2= p_0^2- v_F^2 p_y^2$. One substitutes
(\ref{TM1}) and (\ref{TM2}) into (\ref{SQ}) and uses (\ref{T3}) to
obtain:
\begin{equation}
r_{A_0} = \frac{ p_z(A(p) - p_y^2 v_F^2 B(p)) }{ 2 i \tilde p_z^2 -
p_z (A(p) - p_y^2 v_F^2 B(p)) }
\end{equation}
Since $E_z \sim \partial_z A_0$ the reflection coefficient for the
TM mode is equal
\begin{equation}
r_{TM}= - r_{A_0} = -\frac{ p_z(A(p) - p_y^2 v_F^2 B(p)) }{ 2 i
\tilde p_z^2 - p_z (A(p) - p_y^2 v_F^2 B(p)) } \label{rtm2}
\end{equation}
The reflection coefficients (\ref{rte2}) and (\ref{rtm2}) are
reflection coefficients of transverse electric and transverse
magnetic modes respectively.
One can also rewrite the reflection
coefficients in terms of the polarization tensor components
\begin{equation} r_{\rm TM}=\frac{p_z \Pi_{00}}{p_z \Pi_{00} + 2 i p_y^2},
\qquad r_{\rm TE}= - \frac{ p_z^2 \Pi_{00}+ p_y^2 \Pi_{\rm tr}}
{p_z^2 \Pi_{00} + p_y^2 (\Pi_{\rm tr} - 2 i p_z)}.
\label{rTETM-grPi}
\end{equation}
\section{QED point of view}
The two conditions follow from gauge
invariance:
\begin{align}
p_0 \Pi_{00} (p_0) - p_y \Pi_{y0} (p_0) &= 0 ,\nonumber\\
p_0 \Pi_{0y} (p_0) - p_y \Pi_{yy} (p_0) &= 0 , \nonumber
\end{align}
which yield after Wick rotation
\begin{equation}
p_0^2 \Pi_{00}(i p_0) = - p_y^2 \Pi_{yy} (i p_0) \label{ID1}
\end{equation}
and the property
\begin{equation}
\Pi_{\rm tr} (i p_0) = \Pi_{00} (i p_0) \frac{p_0^2 + p_y^2}{p_y^2}
- \Pi_{xx} (i p_0).
\end{equation}
The reflection coefficients can be rewritten in the form:
\begin{align}
r_{\rm TM}(i p_0) &= - \frac{\sqrt{p_0^2 + p_y^2} \Pi_{yy}(i
p_0)}{2 p_0^2} \Biggl(1 - \frac{\sqrt{p_0^2+ p_y^2} \Pi_{yy}(i
p_0)}{2 p_0^2} \Biggr)^{-1}, \label{Par1} \\
r_{\rm TE}(i p_0) &= \frac{\Pi_{xx}(i
p_0)}{2\sqrt{p_0^2+ p_y^2}} \Biggl(1- \frac{\Pi_{xx}(i
p_0)}{2\sqrt{p_0^2+ p_y^2}}\Biggr)^{-1} .\label{Par2}
\end{align}
In the gauge $A_0=0$ the longitudinal part of the free photon
propagator $D^L$ has the form
\begin{equation}
D^L (ip_0) = \frac{ \sqrt{p_0^2 + p_y^2} e^{-|z|\sqrt{p_0^2 +
p_y^2}} }{2 p_0^2} , \label{DL}
\end{equation}
the transverse part of the free photon propagator $D^T$ has the
form:
\begin{equation}
D^{T} (ip_0) = \frac{e^{-|z|\sqrt{p_0^2+ p_y^2}}}{2\sqrt{p_0^2+
p_y^2}} . \label{DT}
\end{equation}
Let's choose coordinate axes in the plane of a graphene sheet so
that $p_y = {|\bf p|}$. Lifshitz free energy \cite{Lifshitz} has the
form:
\begin{equation}
{\mathcal F}
=T\sum_{n=-\infty}^\infty\int\frac{d^2{\bf p}}{8\pi^2} \ln [(1-e^{-2a\sqrt{\omega_n^2+{\bf p}^2}}r_{\rm
\rm TE}^{(1)}r_{\rm \rm TE}^{(2)})
(1-e^{-2a\sqrt{\omega_n^2+{\bf p}^2}}r_{\rm \rm TM}^{(1)}r_{\rm \rm TM}^{(2)})] ,
\label{EL}
\end{equation}
where $\omega_n=2\pi n T$ are Matsubara frequencies, the respective
transverse magnetic and transverse electric reflection coefficients
from two parallel flat surfaces separated by a vacuum slit $a$ are
denoted by $r_{TM}$ and $r_{TE}$. For the ideal metal $r_{TM}=1$,
$r_{TE}=-1$. Note, however, that some exact results in complicated
geometries\cite{Mar2}-\cite{Mar3} can be essentially different from
the approximations based on Lifshitz formula for two parallel plates
(see also Ref.\cite{Pirozhenko} which considers sphere-plane
system).
From the comparison of the formula (\ref{EL}) and expressions
(\ref{Par1}) -- (\ref{DT}) it follows that for graphene -- ideal
metal, graphene -- graphene the Lifshitz theory takes into account
the set of closed Feynman one loop diagrams responsible for
interaction between the two materials separated by a vacuum slit.
The longitudinal and the transverse parts of the photon propagator
enter two different sets of closed one loop diagrams for the free
energy and multiplied in these Feynman diagrams by the longitudinal
and the transverse components of the polarization operator
$\Pi_{yy}$ and $\Pi_{xx}$ respectively.
Lifshitz type formulas result from the
sum over closed Feynman one loop diagrams with $z=a$ or $z=2a$ in
photon propagators connecting the two sheets of graphene or a
graphene sheet interacting with an ideal metal respectively. Inside
each graphene layer the sum of RPA diagrams is taken into account by
factors in round parentheses in (\ref{Par1}), (\ref{Par2}). The
division of the free energy into longitudinal and transverse parts
in terms of respective parts of photon Green's functions and
polarization operator is equivalent to a division into TM and TE
parts described by the reflection coefficients $r_{\rm TM}$ and
$r_{\rm TE}$ in the Lifshitz approach.
\section{High-temperature asymptotics}
Let's assume $m=0$, $\mu=0$
\cite{Mahan}.
For $|{\bf p}| \to 0$ one gets:
\begin{align}
&\Pi^{00}(ip_0=0) = \frac{4 \alpha N T \ln2}{v_F^2}
+\frac{\alpha N {\bf p}^2}{12 T} + \dots , \nonumber \\
&{\rm tr}\Pi(i p_0=0) - \Pi^{00} (i p_0=0) =
\frac{\alpha N v_F^2 {\bf p}^2}{6T} + \dots , \nonumber
\end{align}
and reflection coefficients have the form:
\begin{align}
r_{TE}(ip_0=0) &\mathop{\simeq} -\frac{\alpha N v_F^2 |{\bf p}|}{\alpha N v_F^2 {|{\bf p}|} + 12 T} , \nonumber\\
r_{TM}(ip_0=0) &\mathop{\simeq} \frac{2\alpha N T \ln2
+ \alpha N {\bf p}^2 v_F^2/(24 T)}{
2\alpha N T\ln2 + \alpha N {\bf p}^2 v_F^2/(24 T) + |{\bf p}| v_F^2}
.\nonumber
\end{align}
Zero Matsubara TM and TE terms yield the following high-temperature
behavior of the free energy (\ref{EL}) in graphene -- ideal metal
system:
\begin{align}
{\mathcal{F}}_{0 \rm TM } &=
-\frac{T\zeta(3)}{16 \pi a^2} + \frac{v_F^2 \zeta(3)}{32\alpha
N\pi (\ln2)a^3} + \dots
, \\
{\mathcal F}_{0\rm TE}
&= -\frac{\alpha N v_F^2 }{192 \pi a^3} + \dots.
\end{align}
Here
\begin{equation}
-\frac{T\zeta(3)}{16 \pi a^2} \equiv {\mathcal F}_{\rm
Drude}\vert_{T\to\infty}=\frac12 {\mathcal F}_{\rm
id}\vert_{T\to\infty} .
\end{equation}
is the high-temperature asymptotics of the metal -- metal system
with a Drude model of the permittivity used
\cite{Sernelius2}-\cite{Drude1}, which is equal to one half of the
high-temperature asymptotics in the metal -- metal system with the
ideal boundary conditions or the plasma model of the permittivity
used \cite{Decca} (see Ref.\cite{Brevik7} for a review). The zero
frequency TE Matsubara term is suppressed by a factor $\alpha N
v_F^2$ and additional power of $1/(Ta)$.
The typical region of validity of the high-temperature asymptotics
for the metal--metal system is $4\pi T a \gg 1$. However, due to a
suppression of nonzero Matsubara terms by the coupling constant
$\alpha$ the free energy in a graphene -- metal system approaches
the high-temperature asymptotics $-T\zeta(3)/(16\pi a^2)$ much
quicker and at shorter separations than in the metal -- metal case
(see next section).
Note that in the graphene -- ideal metal system the high-temperature
asymptotics is derived from the first principles of quantum field
theory.
\section{Nonzero Matsubara terms}
It is often desirable to have an accurate analytical approximation
of the exact result at different separations. We present such an
expression for the sum of nonzero Matsubara terms in this section.
To obtain an appropriate analytical expression we first note that at
separations $H\gg v_F$ one can put $v_F=0$ in any nonzero Matsubara
term. It is possible due to the exponential factor in the Lifshitz
formula which effectively restrains the integration over impulse to
$a |{\bf p}|\lesssim1$. In this case contribution of the type of
$v_F^2 (a |{\bf p}|)^2$ can be neglected compared to $(ap_0)^2=(2
\pi n aT )^2$ due to the smallness of the parameter $v_F$.
In the finite temperature sum of nonzero Matsubara terms in the
Lifshitz formula one can use the reflection coefficients taken at
zero temperature. The corrections due to finite temperature are
suppressed for nonzero Matsubara terms, so we neglect them in the
leading approximation.
Under two mentioned above approximations and the condition $m=\mu=0$
the reflection coefficients of a single graphene layer at zero
temperature have the form:
\begin{align}
r_{\rm TM}^{T=0} &= \frac{\pi \alpha N \sqrt{p_0^2 + {\bf p}^2} }{
\pi
\alpha N \sqrt{p_0^2 + {\bf p}^2} + 8 \sqrt{p_0^2 + v_F^2 {\bf p}^2} }
\simeq \frac{\pi \alpha N \sqrt{p_0^2 + {\bf p}^2} }{ \pi
\alpha N \sqrt{p_0^2 + {\bf p}^2} + 8 |p_0|}
\label{rtm}\\
r_{\rm TE}^{T=0} &= - \frac{\pi \alpha N \sqrt{p_0^2 + v_F^2 {\bf
p}^2} }{
\pi \alpha N \sqrt{p_0^2 + v_F^2 {\bf p}^2} + 8 \sqrt{p_0^2 + {\bf p}^2} }
\simeq - \frac{\pi \alpha N |p_0|}
{\pi \alpha N |p_0| + 8 \sqrt{p_0^2 + {\bf p}^2} }
\label{rte}
\end{align}
\begin{figure}
\centering \includegraphics[width=12cm]{plotmay.eps} \caption{Ratio
$\rho_1$ of the free energy to the leading high-temperature
asymptotics $-T\zeta(3)/(16\pi a^2)$. Both graphs are evaluated for
$T=300$K. In graphene the values $m=\mu=0$ were used.}
\label{Mayplot}
\end{figure}
Due to smallness of the reflection coefficients (both being of the
order of $\alpha$) we can take just the first term in the expansion
of the logarithm in the Lifshitz formula. Note, however, that
expansion of the TM reflection coefficient in $\alpha$ is not
legitimate, as will become evident below (see (\ref{appt})).
\begin{figure}
\centering \includegraphics[width=12cm]{TMTE4.eps} \caption{Ratio
$\rho_2$ of the free energy for a graphene - metal system with
$\mu=m=0$ to the ideal metal - ideal metal free energy at $T=300$K.}
\label{ratio1}
\end{figure}
The sum of nonzero Matsubara terms in (\ref{EL}) in this
approximation in the TM case with $r^{(1)}_{\rm TM} = r_{\rm
TM}^{T=0}$ (\ref{rtm}) and $r^{(2)}_{\rm TM}=1$ equals to
\begin{align}
\Delta\mathcal{F}_{\rm TM} &= - \frac{T}{2\pi} \sum_{n=1}^{+\infty}
\int_{Hn/2}^{+\infty} ds_1 \frac{s_1^2}{s_1+ 16nT/(\alpha N)}
\exp(-2as_1) = \nonumber \\
&= - \frac{T}{8\pi a^2} \sum_{n=1}^{+\infty} \exp(-Hn) (1 - gn + Hn)
+ (gn)^2 \exp(gn) {\rm E_1}(gn + Hn) , \label{TMint2}
\end{align}
here $s_1=\sqrt{\omega_n^2+ |{\bf p}|^2}$ and $g \equiv 32 T
a/(\alpha N)$, $E_1$ stands for the standard exponential integral
function. It is convenient to reexpress the result (\ref{TMint2}) in
an integral form. For this purpose one has to differentiate
(\ref{TMint2}) over $H$, assuming $H$ as an independent parameter
for the moment, calculate the sum over $n$ and then integrate back
over $H$ (the integration constant is fixed as zero at
$H\to\infty$). Thus one obtains
\begin{equation}
\Delta\mathcal{F}_{\rm TM} = -\frac{T \alpha N}{8 a^2}
\int_{H}^{+\infty} dt \frac{\exp(t)\bigl(\exp(t)+1 \bigr)\:
t^2}{\bigl(\exp(t)-1 \bigr)^3 \bigl(8H + \pi\alpha N t \bigr)}
\label{Intrepr}
\end{equation}
The TE part of the nonzero Matsubara terms of the Lifshitz formula
with the coefficients $r^{(1)}_{TE} = r_{TE}^{T=0}$ from (\ref{rte})
and $r^{(2)}_{TE}= -1$ gives the following contribution in the
leading order in $\alpha$
\begin{equation}
\Delta\mathcal{F}_{TE} \simeq - \frac{T^2 \pi \alpha N }{8}
\sum_{n=1}^{+\infty} n \int_{\omega_n}^{+\infty} ds_1 \: \exp(-2 a s_1)
= - \frac{T^2 \pi \alpha N }{16 a} \frac{\exp(-H)}{(1-\exp(-H))^2} .
\label{TE1}
\end{equation}
Thus, the complete result for the sum of nonzero Matsubara TM and TE
terms in the approximation described above is given by
(\ref{Intrepr}) and (\ref{TE1}):
\begin{equation}
\Delta\mathcal{F} = \Delta\mathcal{F}_{TM} +\Delta\mathcal{F}_{TE}.
\label{nonz}
\end{equation}
Consequently, the leading $v_F=0$ contribution to the free energy is
the sum $-T\zeta(3)/(16\pi a^2)$ + $\Delta \mathcal{F}$. It can be
used for the comparison of the theory and experiment with $1\%$
accuracy for all separations at $T=300$K.
From (\ref{Intrepr}) and (\ref{TE1}) one gets the energy at $T=0$ in
the limit $v_F\to 0$:
\begin{equation}
\Delta\mathcal{F}\bigl|_{T\to 0} =
-\frac{\alpha N}{128\pi a^3} \ln\bigl(1+ 8/(\alpha N \pi)\bigr)
- \frac{\alpha N}{256 \pi a^3 } , \label{appt}
\end{equation}
where the non analyticity in $\alpha$ comes from the TM mode.
Fig.\ref{Mayplot} clearly demonstrates that the free energy of a
graphene -- metal system approaches the high-temperature asymptotics
$-T\zeta(3)/(16\pi a^2)$ much quicker and at shorter separations
than the two metals' system. Such an approach to the
high-temperature behavior is related to the fact that nonzero
Matsubara terms are of the order of the coupling constant $\alpha$
and thus very small in comparison with respective terms for metals.
The zero frequency TM
Matsubara term acquires the value of the sum of nonzero Matsubara
terms in the free energy at separations $a\approx 100$nm at $T=300$K
(see Fig.\ref{ratio1}) and dominates in the free energy at larger
separations. Thus the high-temperature behavior in graphene --
metal systems should be observed at separations of the order of
$100$ nm at $T=300$K (the same effect in metal -- metal systems
takes place at separations of the order of several micrometers at
$T=300$K).
\section{Conclusions}
The behavior of the free energy of the graphene -- metal system is
studied on the basis of the field theoretic model. The components of
the polarization operator of $2+1$ quasiparticles in a graphene
layer are evaluated at finite temperature. The TM and TE reflection
coefficients are derived from the solutions of the boundary problems
for vector potentials.
In the high-temperature limit the asymptotics of the free energy
coincides with the Drude model asymptotics for two metals' system.
The crossover to the high-temperature behavior in a graphene-metal
system takes place at separations $a$ of the order of $100$ nm at
$T=300$K. This is the reason why the systems with graphene are very
promising for the experimental studies of the finite-temperature
Casimir effect.
\section*{Acknowledgments}
The author is grateful to the organizers of QFEXT-11 for support.
The author is grateful to colleagues for numerous discussions in
Benasque.
|
\section{Introduction}
Whereas the gravitational nature of Dark Matter (DM) is a crucial
ingredient for the success of the standard cosmological model, its
non-gravitational character proves elusive and remains essentially
unknown to date. The possible connection with the electroweak scale
and the associated expectation of an interaction strength not much
smaller than $G_F$ and a DM mass not much lighter than
$\mathcal{O}(10\,\mathrm{GeV})$ has motivated the construction of
experiments looking for the direct scattering of DM against nuclei in
the lab~\cite{witten}. The experimental efforts in the field are
proceeding in full force and at least another decade of progress is
expected.
With the main observable being the nuclear recoil spectrum the
information content is rather limited. Moreover, as was realized over
the past decade, the recoil spectrum resulting from DM collisions is
model dependent and may not follow the simple exponential rise towards
low energies expected from elastic scattering~\cite{iDM,CiDM, MDDM, MiDM,
FFDM}. On the other hand, one of the most robust predictions of the
cold DM model is that the relative velocity of the Earth against the
DM halo should vary with the Earth revolution around the
Sun~\cite{Drukier:1986tm, Freese}. The expected recoil rate is therefore a general
periodic function with a fundamental period of one year with a
particular phase. Despite the strong motivations to look for such
modulations, it has so far been achieved by only two experiments,
DAMA~\cite{Bernabei:2008yi,Bernabei:2010mq} and
CoGeNT~\cite{Aalseth:2011wp}. The difficulty is of course in
maintaining the stability of the apparatus and the rejection of
background over a long period of time. Indeed, the two experiments
just mentioned have experienced considerably more background events
than some of the other extremely clean experiments such as
\mbox{CDMS-II}~\cite{cdms,Ahmed:2010wy} and
XENON100~\cite{xenon100}. Moreover, the search for annual modulations
is not without difficulties since many more mundane phenomena are
known to exhibit such modulations. Therefore, two of the main goals of
this paper are 1) to show that the time series of the reported DAMA
and CoGeNT signals can be used to make firm statements about
background hypotheses proposed in the literature and 2) to establish
further observables based on the time distribution of the energy
spectrum.
The only direct detection experiment which claims detection of a firm
signal from dark matter is DAMA, situated in the underground LNGS
laboratory at Gran Sasso, Italy. The DAMA dataset consists of two main
periods, DAMA/NaI (Dec 1995 - July 2002) and DAMA/LIBRA (Sept 2003 -
Sept 2009), amounting to a cumulative exposure of
1.17~ton$\times$years. The residual rate shown by the collaboration
exhibits a very clear annual modulation compatible with what is
expected from DM models where the Earth's motion around the Sun
results in a modulation peaking on approximately June 2$^{\rm
nd}$. The collaboration has also released the modulation amplitude
as a function of recoil energy, which seems to exhibit a peculiar form
possibly more consistent with inelastic scattering~\cite{iDM,CiDM,MDDM,
FFDM,MiDM}. It should be kept in mind, however, that this energy
spectrum is obtained from the full data set by assuming the periodic
functional form. Sadly, there has been no release of the full
data set to date. It is particularly unfortunate because, as we will
see below, the procedure carried by the collaboration to obtain the
power spectrum of the signal and other parameters makes it difficult
to compare to both background and signal hypotheses.
Given the seasonal nature of the signal it is entirely conceivable
that environmental effects induce backgrounds with an annual
modulation pattern just like the one seen by DAMA. As such, it is
clear that such contamination may depart significantly from a
sinusoidal distribution in time. It is therefore important---whenever
enough data on a putative background inducing process is
available---to assess its viability based on a full time-series
analysis. For example, in this work we employ Pearson's coefficient of
correlation as a test statistic when quantifying the compatibility of
two data sets. This allows for a quantitative comparison between two
data sets which is independent of any assumption about the functional
form of the time variations.
Fluxes of neutrons probably constitute the most dangerous source of
background as they lead to nuclear recoils which can mimic DM-nucleus
interactions \cite{ralston}. Indeed, it is known from ICARUS
measurements at LNGS \cite{icarus} that the ambient neutron flux
generated in the surrounding rock shows some variation on the
timescale of a year. However, with a total of five data points we find
that it is not possible to make any \textit{statistically} significant
statement regarding the temporal compatibility of this candidate
background with the DAMA signal. We therefore choose not to elaborate
on this issue any further.
In contrast to rock-generated neutrons, a wealth of data is available
on another guaranteed source of neutrons: cosmic ray muons which can
penetrate deep underground and induce spallation reactions in the
detector and nearby. It is also possible for these muons to deposit
their energy directly into the detector crystals~\cite{nygren}.
Measurements of the muon flux underground have a long history and its
seasonal variation is firmly established. There are published
measurements at LNGS from MACRO~\cite{Ambrosio:1997tc},
LVD~\cite{Selvi} and most recently from
Borexino~\cite{D'Angelo:2011fs}. For DAMA, the relevant measurement is
the one from LVD since it was taken at the same underground lab and it
overlaps in time with the DAMA/LIBRA runs 1--5. Figure~\ref{fig:lngs}
shows the percent residuals of the muon flux when binned in
concordance with DAMA with the annual mean count rate subtracted. Also
shown are the residuals in the 2--4\,keV bin of the DAMA/LIBRA data
assuming a baseline rate of $\bar s = 1.15$~cpd/kg/keV. The seeming
similarity in time and amplitude is tantalizing and it has lead
various authors to put forward the hypothesis that both signals are in
fact measurements of the very same cosmic ray
phenomenon~\cite{ralston,nygren,Blum:2011jf}. It will be one of the
central points of the paper to critically examine these claims,
finding that they have difficulty explaining the observed modulations.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=\columnwidth]{TS_DAMA+MUONSresiduals.pdf}
\end{center}
\caption{Percent annual residuals of the LVD measured muon flux when
binned in accordance with the DAMA/LIBRA runs~1--5. The latter
residuals are shown for the 2--4\,keV bins assuming a baseline count
rate of $\bar s = 1.15$~cpd/kg/keV.}
\label{fig:lngs}
\end{figure}
The issue of annual modulation in direct DM detection has recently
received further impetus from the CoGeNT experiment, located in the
Soudan Mine, MN, USA. The collaboration has published its analysis on
a 458 day run with a 440\,gram Ge detector
\cite{Aalseth:2011wp}. After removal of cosmogenic background
contamination, the data exhibits a seasonal variation peaking around
the middle of April. As we shall see below the modulation is manifest
only in the higher energy part of the recoil spectrum where the energy
spectrum is rather flat with none of the usual features expected from
the direct detection signal of DM, elastic or otherwise. The
unexplained exponential rise in the recoil spectrum at lower energies,
which has stirred quite a commotion in the recent past, shows no
evidence of annual modulations. Making use of the time-stamped data
provided by the collaboration we address the modulation part of the
signal (see also ref.~\cite{Fox:2011px,volansky} for prior analyses)
and investigate the potential role of cosmic ray muons for the CoGeNT
detector.
In light of the significant advances in sensitivity from direct
detection experiments achieved in the past decade and future
improvements expected in the next one, we also address the question of
whether the time-series of a signal may encode additional evidence
that it is DM which is scattering on a target. In particular we point
out that higher harmonics may be present in an annually modulated
signal. We show how this signature manifests itself in the scattering
rate and explore to some extent its sensitivity on the assumed halo
parameters.
The paper is organized as follows: We start with
section~\ref{sec:null} by establishing statistical evidence for
periodic variation in the data sets under consideration. Restricting
ourselves to a sinusoidal form of the signals, in
section~\ref{sec:phase} we allow both phase and period to vary and
explore the inferred values of these two parameters from the different
datasets. In section~\ref{sec:correlation} we drop the assumption of
a sinusoidal form and instead directly compare the time series of the
various data sets by performing a correlation
analysis. Section~\ref{sec:biannual} discusses the effect of higher
harmonic modulations as a new diagnostic tool when searching for dark
matter in the lab. We summarize the main conclusions and findings of
the paper in section~\ref{sec:conclusions}.
\section{Rejecting the Null Hypothesis}
\label{sec:null}
The existence of a periodic signal in the DAMA and LVD data sets is
clear even without any sophisticated statistical analysis. The
situation is somewhat more subtle in the case of CoGeNT. Both for the
purpose of completeness, and because it reveals interesting
differences between the data sets, we begin our analysis by testing
the null hypothesis of no signal (\textit{i.e.}~only pure noise) in
each of the data sets. The most convenient way of doing such a
spectral analysis is by looking at a periodogram of the data.
The classical periodogram of Schuster~\cite{schuster1890} allows for a
calculation of the power spectrum of a signal even in the case of
uneven sampling of the data. In ref.~\cite{1982ApJ...263..835S}
Scargle showed that a modified version of the classical periodogram,
which we call the Lomb-Scargle (LS) power spectrum (and sometimes just
power spectrum), results in the same well defined statistical behavior
as the Fourier power spectrum used in the case of even sampling. The
LS power spectrum allows one to reject noise at a $1-\alpha$
confidence level for a single, preselected frequency by demanding a
power level in excess of $z_\alpha = \ln\left(\alpha^{-1}\right)$. If
one is testing $N$ independent frequencies then the power level
required increases to $z_\alpha =-\ln \left[ 1-(1-\alpha)^{1/N}
\right]$ which contains the statistical penalty for inspecting more
than one frequency. This means that when testing for the null
hypothesis over a range of frequencies we must employ some care in
choosing the frequencies to be tested if we want a meaningful
statistical interpretation. If the time series is evenly distributed
then the standard choice for the frequencies in the discrete Fourier
transform results in independent frequencies. Since the data sets we
consider are not grossly uneven and approximately cover an integer
number of years we chose $\omega_n = n \omega_F$ with the fundamental
frequency $\omega_F = 2\pi/T$ where $T$ is roughly the range of years
covered by the data set\footnote{The power spectra in this section all
display a peak at a period of exactly one year. We caution the
reader that this does not imply the best fit value would be exactly
one year, but is simply an artifact of the frequencies we chose to
sample when looking to reject the null hypothesis.}. We checked that this results in low correlation among the
test frequencies using the procedure described in App.~D of
Ref.~\cite{1982ApJ...263..835S}.
\subsection{DAMA and LVD data}
We start by considering the measurements of the muon flux by the LVD
experiment~\cite{Selvi}. The average integral muon intensity
underground is reported as $\langle I_{\mu} \rangle \simeq 3\times
10^{-4}\,\mathrm{m}^{-2}\,\sec^{-1}$ with an annual variation of $\sim
2\%$ in amplitude as can be seen from Fig.~\ref{fig:lngs}. The data
spans a total of eight years with more than 2800 data points which we
obtain by digitizing Fig.~2 of~\cite{Selvi}.
The solid line in Fig.~\ref{fig:LVD_null} shows the power spectrum
obtained from the full LVD data set. Aside from the yearly modulation,
which is plainly visible from the time series itself, the power
spectrum also speaks unequivocally of the existence of temporal
variations on time-scales greater than a year. That same figure
(dashed line) shows the effect of subtracting the mean intensity from
the data on a \textit{yearly} basis as done by the DAMA collaboration
with their own data. The figure makes it clear that such a procedure
would mask most of the power at periods much greater than a
year. Nevertheless, we note that substantial power remains in modes
with periods a little over one year.
Figure~\ref{fig:DAMA-LIBRA_null} shows the power spectrum for the
DAMA/LIBRA $2-4~{\rm keV_{ee}}$ data set. In contrast to the LVD spectrum,
there is little power at periods greater than a year. This, however,
may simply be an artifact of the way the DAMA/LIBRA modulation data is
obtained. The DAMA collaboration calculates the residuals by
subtracting the mean on a yearly basis for each cycle. As mentioned in
the previous paragraph such a procedure tends to dampen power at
periods much greater than a year. Nevertheless, the absence of any
power even for periods slightly over a year, as seen in the LVD data
above, is interesting and serves as the first distinguishing feature
between the two data sets. Unfortunately, it is difficult to assign a
quantitative significance to this difference without the full,
unsubtracted time series. This issue further motivates the release of
the unsubtracted data by the DAMA collaboration to allow for a proper
comparison of the power spectrum. Finally, we note that no biannual
mode (T = 1/2 year) is present in the DAMA/LIBRA power spectrum. We
will comment on this further in Sec.~\ref{sec:biannual} below.
\begin{figure}[t]
\begin{center}
\includegraphics[width=\columnwidth]{LVD_null.pdf}
\end{center}
\caption{The Lomb-Scargle power spectrum of the LVD data as a function
of the period (solid-blue). Also drawn are the 99\% confidence lines
for excluding the noise hypothesis for a single frequency (lower) or
any of the frequencies shown (top). Note the substantial power
present for modes with a period greater than a year. The dashed-red
curve shows the effect of subtracting the yearly mean from the data
on a yearly basis, as is done by the DAMA collaboration. The
subtraction has little effect on the high frequency modes, but
results in a strong damping of the long period modes as can be
expected.}
\label{fig:LVD_null}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=\columnwidth]{DAMA-LIBRA_null.pdf}
\end{center}
\caption{The Lomb-Scargle power spectrum of the DAMA-LIBRA data in the
$2-4~{\rm keV_{ee}}$ energy region as a function of the period (solid-blue).}
\label{fig:DAMA-LIBRA_null}
\end{figure}
\subsection{CoGeNT data}
The CoGeNT collaboration has released the time-stamped data of their
442~live-day run~\cite{Aalseth:2011wp} for public use. This makes the
computation of the LS diagram in principle straight-forward. However,
the data also suffers from background activity of electron capture
decays of cosmogenically activated long-lived isotopes. A measurement
of higher energetic K-shell captures together with reported ratios of
L- to K-shell decays allows one to correct for the L-shell activity in
the DM acceptance region.
Based on the expected cosmogenic activity in the CoGeNT dataset it
seems reasonable to divide the low energy data into three regions:
$0.5-0.9~{\rm keV_{ee}}$, $0.9-1.6~{\rm keV_{ee}}$, and $1.6-3.0~{\rm keV_{ee}}$. Considerable
cosmogenic activity is observed in the middle region. In contrast only
a small amount of cosmogenic activity is expected in the low region
($0.5-0.9~{\rm keV_{ee}}$) and very little if any is expected in the high
region ($1.6-3.0~{\rm keV_{ee}}$). Since the participating isotopes are rather
long-lived ($t_{1/2}\gtrsim 200\,\mathrm{days}$) they are expected to
result in substantial power at long periods in the corresponding
Fourier power spectrum. We have verified that this is indeed the
case. However, since we are interested in more localized phenomena in
frequency space, such as annual modulations, we need to subtract the
cosmogenic activity from the data. We have done this on a daily basis
based on the reported activity levels~\cite{Juan}. We have verified
that the results remain qualitatively the same even when adopting a
different subtraction procedure, where we fit for the exponentially
decaying component.
Figure~\ref{fig:CoGeNT_null} shows the power spectrum for the subtracted
data in the three energy regions. The power spectrum was calculated in
30 equally spaced frequencies with a fundamental frequency of 2
years. The null hypothesis of noise can only be confidently rejected
for the $1.6-3.0~{\rm keV_{ee}}$ region, where considerable power is present at
a period of about one year. However, the same cannot be said about the
lower energy regions where no significant power is observed. Thus, one
can claim the detection of annual modulations at about $99\%$
confidence level only in the $1.6-3.0~{\rm keV_{ee}}$ region.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\columnwidth]{CoGeNT_null.pdf}
\end{center}
\caption{The Lomb-Scargle power spectrum of the subtracted CoGeNT data
as a function of the period for the three energy regions:
$1.6-3.0~{\rm keV_{ee}}$ (solid-black), $0.9-1.6~{\rm keV_{ee}}$ (dashed-blue), and
$0.5-0.9~{\rm keV_{ee}}$ (dotted-purple).}
\label{fig:CoGeNT_null}
\end{figure}
\section{Phase Analysis}
\label{sec:phase}
Aside from the period, the second most important characteristic of the
oscillations observed by the DAMA experiment is the phase of the
signal. Dark matter collision rate with the detector is expected to
peak on June 2$^{\rm nd}$, corresponding to $t_0 =152$ days after
January 1$^{\rm st}$. In this section we investigate the phase
associated with the oscillations seen in the DAMA and CoGeNT data and
compare them to the phase of the muon intensity modulations. This
comparison was already attempted by the DAMA collaboration itself,
however, it was criticized by refs.~\cite{nygren,Blum:2011jf} on two
accounts. First, in their fit to the data the DAMA collaboration fix
the period and allow only the phase to float. Second, the underlying
signal may not be truly sinusoidal which may invalidate the
statistical inference of a fit to a sine function. We address the
first issue in this section and investigate the second problem in the
next.
\subsection{Frequentist Approach}
\begin{figure}[tb]
\begin{center}
\includegraphics[width=\columnwidth]{TS_DAMAt0omega.pdf}
\end{center}
\caption{Confidence regions in period $T$ and phase $t_0$ obtained
from a $\chi^2$ fit to the DAMA/LIBRA residuals. The various lines
as labeled correspond to the three different energy binnings
provided by the DAMA collaboration with the time origin set to Jan
1, 2003. The shaded region illustrates the effect of shifting the
time origin to Jan 1, 2011 for the 2--4\,keV residuals and the light
dot-dashed ellipse shows the time origin shift to Jan 1, 1995.}
\label{fig:freq}
\end{figure}
We start with a simple frequentist approach and investigate the effect
of departing from strict annual periodicity (i.e $T$ is
not necessarily 365 days) on the DAMA phase $t_0$. Under the premise
of a sinusoidal signal, $A\times\cos[\omega(t-t_0)]$ with
$\omega=2\pi/T$, we fit amplitude $A$ to the DAMA/LIBRA residuals by
minimizing the usual $\chi^2$ function while scanning over period $T$
and phase $t_0$. A confidence region in $t_0$ and $T$ can be
constructed from the profile likelihood method~\cite{pdg2010} which
effectively removes the dependence on the nuisance parameter $A$. This
way, we first obtain the global minimum $\chi^2_{\mathrm{min}} $ from
which the confidence region is obtained by requiring
\begin{equation}
\label{eq:conf-reg}
\chi^2(\hat A,t_0,T) \leq \chi^2_{\mathrm{min}} + \Delta \chi^2 ,
\end{equation}
where $\hat A$ is the maximum likelihood estimate for $A$ at each
point $(t_0,T)$. For a coverage probability of 95\% one chooses
$\Delta \chi^2 = 5.99$. Figure~\ref{fig:freq} shows the result of
fitting the DAMA/LIBRA residuals in the $(T,t_0)$-plane. The three
ellipses give the 95\%~C.L. regions for the various energy binnings as
provided by the DAMA collaboration. As can be seen the data---at the
required confidence---are not necessarily consistent with the dark
matter interpretation ($t_0=152$\,days and $T=1\,$yr as indicated by
the thin gray lines). As we argue in the next paragraph, the
interpretation of such fits have to be handled with some care.
When allowing the period to float it is important to realize that
statements about the inferred phase become dependent on the starting
date. First, there is the obvious effect that the phase is measured in
days with respect to Jan~1$^{\rm st}$, but if the period is not one
full year, then which year is used as the origin affects the
phase. Comparison of the phase in days between experiments with
different origins is meaningless until the origins are made to
coincide. This effect is simple to correct for and requires that we
use the same time origin for the different data sets. In
Fig.~\ref{fig:freq} the chosen time-origin for the three ellipses
showing the various energy bins is Jan 1, 2003.
There is another, more subtle issue that affects the determination of
the phase when the period is allowed to float. As can be seen from the
ellipses in Fig.~\ref{fig:freq} the DAMA data exhibits an
anti-correlation between the period and the phase. Fits with periods
longer than a year strongly favour phases smaller than $t_0\sim 150$
days and vice-versa. Note, however, that the sign of the correlation
itself depends on the chosen origin. When shifting the latter forward
such that the DAMA/LIBRA data lies in the past, $t_0$ and $T$ become
positively correlated instead. This is illustrated in
Fig.~\ref{fig:freq} by the filled ellipse obtained from the 2--4\,keV
energy bins for which we have chosen Jan 1, 2011 as the origin.
It is important to note that the distribution in $t_0$,
given a certain period, depends on the chosen time origin. Ideally,
we would like to make coordinate-independent statements and care must
be taken when interpreting the results. This issue further motivates
the correlation study provided in the next section.
Given the above, before proceeding and comparing the LVD and DAMA data sets we must decide on a common time origin. The starting date used by the
DAMA collaboration (Jan 1$^{\rm st}$, 1995) is six years apart from
the LVD data (Jan 1$^{\rm st}$, 2001). In our analysis we will
concentrate on the DAMA/LIBRA data, which started on September 9$^{\rm
th}$, 2003, and so a sensible choice is January 1$^{\rm st}$, 2003
as the origin since it has sufficient overlap with both data sets.
To quantify the level of agreement of the muon-induced background
hypothesis with DAMA we switch now to a Bayesian approach. This allows for a convenient generalization of the Lomb-Scargle periodogram into the two dimensional phase-frequency space as discussed in Appendix~\ref{sec:bayes-vers-lomb} . Such an
approach is particularly convenient when one wishes to incorporate
further assumptions on the provided data sets (e.g. choosing priors on
the period, phase, and amplitude). We will not incorporate any such
assumptions in the current analysis, since we would like to keep it as
unbiased as possible. Moreover, we have verified that the conclusions
arrived at below remain the same when we employ a frequentist analysis
instead.
\subsection{Bayesian Approach}
The Bayesian approach can conveniently generalize the Lomb-Scargle periodogram and allow for the testing of different hypotheses and priors. In this section we
restrict ourselves to modelling the data with a simple sinusoidal
function $f(t)=A\times \cos{[\omega(t-t_0)]}$ but following through
the steps presented in Appendix~\ref{sec:bayes-vers-lomb}, a
generalization to more complicated test-functions is in principle
straightforward.
We are mainly interested in the relationship between phase $t_0$ and
period $T$. The posterior probability $P(\{\omega,t_0 \}||d)$,
\textit{i.e.} the probability of observing frequency $\omega =
2\pi/T$, and phase $ t_0$ given the data $d$ is obtained by
integrating the full posterior $P(\{A,\omega,t_0 \}||d)$ over the
amplitude $A$ restricted to positive values\footnote{It is possible to
obtain a more compact expression for the posterior without the error
function by integrating over all amplitude values both positive and
negative. However, that leads to a $\pm\pi$ ambiguity in the
phase. While this degeneracy is easily removable by inspection, we
prefer to avoid this complication in what follows.},
\begin{eqnarray}
P(\{\omega,t_0 \}||d) &\propto& \frac{\sigma^{1-N}}{\sqrt{p}} \left[1+{\rm Erf}\left(\frac{h}{\sqrt{2\sigma^2 p}} \right) \right] \nonumber \\ &\times& \exp\left(\frac{h^2}{2 \sigma^2 p}\right) ,
\label{eq:posterior}
\end{eqnarray}
where,
\begin{eqnarray}
p&=& \sum_i \cos^2\omega\left(t_i-t_0\right) ,\\
h&=& \sum_i d_i \cos\omega\left(t_i-t_0\right) .
\end{eqnarray}
Before proceeding and using~(\ref{eq:posterior}) in the full
two-dimensional period-phase space we can make contact with the
procedure which is usually employed, \textit{i.e.} fixing the period
to one year, $T=365$~days. Thereby we impose a delta function prior on the
period centered at one year from which we consequently obtain the
posterior probability in phase only, $P(\{t_0 \}||d)$. We find that
the DAMA/LIBRA and LVD data do not agree with respective
values
\begin{align}
\label{eq:phase-posterior}
t_0(\mathrm{DAMA}) & = (131\pm 13)\,\mathrm{days} ,\\
t_0(\mathrm{LVD}) & = (187\pm 2)~\mathrm{days} .
\end{align}
The tight error bars on the LVD data may at first seem surprising. To
better motivate this number we consider that given the 8 years of
data, one can easily determine by eye a phase shift of about 20
days. Statistics allows a further reduction of about an order of
magnitude because of the very large number of points available in the
dataset. However, it should be kept in mind that the uncertainty
quoted is only statistical and does not take into account the possible
systematic issues associated with the obvious presence of temporal
variations on larger time scales\footnote{Moreover, there is a systematic error associated with the digitization of the data that amounts to an additional uncertainty of a couple of days.}. In fact, using the error bars of the
LVD data, the resulting $\chi^2$ of a fit to a sinusoidal function is
extremely poor. Hence, the numbers quoted above should really only be
understood as the best inference on the underlying parameters of a
model consisting of a single frequency.
To see if the discrepancy between DAMA/LIBRA and LVD remains when
relaxing the assumption of a strict annual periodicity, we now
evaluate~(\ref{eq:posterior}) for the two data sets. The result is
shown in Fig.~\ref{fig:PeriodPhaseComparison}. This figure clearly
shows that the inferred modulations seen in the two data sets---if
interpreted as sinusoidal modulations---are incompatible with each
other.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\columnwidth]{PeriodPhaseComparison.pdf}
\end{center}
\caption{A comparison of the period-phase posterior for DAMA/LIBRA
(bottom) and the LVD phase (top). Shown are the 68\%, 90\%, and 99\%
credibility ellipses.}
\label{fig:PeriodPhaseComparison}
\end{figure}
A similar analysis can be done for the CoGeNT data%
\footnote{A Bayesian approach to the CoGeNT data has very recently
also been taken by~\cite{Arina:2011zh}.},
as shown in Fig.~\ref{fig:CoGeNTPeriodPhase}, where we chose to use
the entire low energy range ($0.5-3.0~{\rm keV_{ee}}$). As discussed in the
previous section, evidence for modulation is only present in the
restricted range of $1.6-3.0~{\rm keV_{ee}}$. However, isolated power on an
annual time scale is present in the entire range and it is therefore
not unreasonable to employ the full range when looking to make
inferences about the modulation parameters. CoGeNT's credibility
ellipses are compared with those obtained from the MINOS
data~\cite{Adamson:2009zf}. From there, we arrive at similar
conclusions as before, namely, that the muon and direct detection data
sets seem incompatible We have verified that when considering the
restricted range $1.6-3.0~{\rm keV_{ee}}$ in the CoGeNT data we arrive at the
same conclusions. We note in passing that the orientation of the
MINOS ellipses---which are obtained with a time origin of Jan
$1^{\mathrm{st}}$, 2010---is similar to the one found in
Fig.~\ref{fig:freq} for DAMA once one shifts the time origin to the
future of the data set.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\columnwidth]{CoGeNTPeriodPhase.pdf}
\end{center}
\caption{The period-phase posterior for the full low energy CoGeNT
data ($0.5-3.0~{\rm keV_{ee}}$)~\cite{Aalseth:2011wp} and MINOS
data~\cite{Adamson:2009zf}. Shown are the 68\%, and 90\% credibility
ellipses. The time origin in both cases was chosen to be January
1$^\text{st}$, 2010. }
\label{fig:CoGeNTPeriodPhase}
\end{figure}
The credibility contours in Fig.~\ref{fig:CoGeNTPeriodPhase} indicate
a range of viable parameters which is almost twice as large as what
has been quoted in the CoGeNT release paper~\cite{Aalseth:2011wp}.
For a direct comparison we thus also perform a frequentist fit to a
cosine function (plus a constant.) For example, when using the Minuit
package~\cite{minuit} for the $\chi^2$-minimization, we obtain $T =350
\pm 26$ days, $t_0 = 110\pm 13$ days, and $A = 17\%\pm 4\%$ for a time
binning resulting in 21~bins. This is in good agreement
with~\cite{Aalseth:2011wp}. However, it should be stressed that the
quoted errors are obtained from a default value of $\Delta \chi^2=1$
which does not reflect the increased freedom of fitting more than one
parameter. Furthermore $\chi^2 $ does not grow very large with respect to
its rather small minimum value, $\chi^2_{\mathrm{min}}/d.o.f. =
6.8/17$ and even a fit to a constant rate yields a reasonable
$\chi^2_{\mathrm{min}}/d.o.f. = 26.9/20$. This can be traced back to
large error bars in conjunction with limited data. As a result of
this, contours with nominally larger confidence will rapidly open
up the parameter space. This, however, does not signal real
compatibility of the data sets because of the aforementioned reasons.
\section{Correlation Analysis}
\label{sec:correlation}
A valid concern with regard to the analysis above is the underlying
assumption of sinusoidal variation. The power-spectrum of the LVD
data, Fig.~\ref{fig:LVD_null}, makes it clear that power exists in
modes with periods larger as well as smaller than one year. One may
then worry that the phase comparison discussed in the previous section
suffers from a systematic misinterpretation. We do not entirely
endorse this concern because it is difficult to understand how the
very prominent annual modulation in the LVD data, with its very
definite phase, can be masked by the much weaker modulations at other
frequencies. Nevertheless, we now set to investigate this issue in a
way that does not rely on assuming any particular functional form for
the underlying temporal variations.
Moreover, even if the timing between muons and DAMA (and possibly
CoGeNT) are incommensurate at first sight, could it be that the
underlying stochastic nature of the background process alleviates the
observed tension in the annual phase? This explanation for DAMA was
suggested recently in an interesting paper by Blum~\cite{Blum:2011jf}.
Assuming a simple, generic model for how a muon-sourced background may
be realized, the distribution in phase originating from the Poissonian
process is claimed to be compatible with the one observed by DAMA.
In this section we address the above issues by performing a
correlation analysis. In particular we follow Ref.~\cite{Blum:2011jf}
and generate mock data for DAMA based on the muon hypothesis. We show
that the resulting mock data exhibits a substantially higher degree of
correlation with the actual muon data than does the real data from
DAMA thus ruling out the hypothesis. We then go on and use a similar
correlation analysis to show that the muon hypothesis is also an
unlikely cause for the modulations seen by CoGeNT.
\subsection{DAMA}
\label{sec:dama-corr}
Since there is considerable overlap between the DAMA/LIBRA and LVD
data, it is straightforward to define Pearson's coefficient of
correlation. We first bin the LVD data according to the DAMA/LIBRA
bins. Then the sample's correlation coefficient is defined as,
\begin{eqnarray}
\label{eq:pearsonr}
r_{X,Y} = \frac{1}{n-1}\frac{\sum_{i=1}^{n}\left(X_i - \bar{X}\right)\left(Y_i - \bar{Y}\right) }{\sigma_X\sigma_Y}
\end{eqnarray}
where $n$ is the number of overlap bins, $\bar{X}$ and $\bar{Y}$ are
the samples' mean, and $\sigma_{X,Y}$ are the samples' standard
deviations. The first statistical question we address is whether we
can exclude the no-correlation hypothesis. The answer to this question
lies in the statistic,
\begin{eqnarray}
t = \frac{r\sqrt{n-2}}{\sqrt{1-r^2}}
\end{eqnarray}
which---in the case of the null-hypothesis---is known to follow the
Student's distribution with $d.o.f = n-2$. The DAMA/LIBRA and LVD
data share 39 temporal bins and so the 90\% (99\%) two-sided exclusion
limit on the statistic is $|t| > 1.687 (2.715)$.
For the correlation between the two data sets we find a value,
\begin{equation}
\label{eq:r}
r_{\mathrm{LVD,DAMA}} = 0.44\quad \Rightarrow\quad t_{\mathrm{LVD,DAMA}} = 2.95 ,
\end{equation}
which allows us to exclude the no-correlation hypothesis with
confidence level greater than 99\%. This, however, should come as no
surprise since both datasets exhibit strong annual modulations with a
similar phase. Any such samples will exhibit a correlation at some
level, but that of course does not imply that they are indeed causally
related. The more interesting question we would like to address now
is whether we can exclude causation. This question can only be
answered in a model-dependent way.
The model we consider was presented by Blum in
Ref.~\cite{Blum:2011jf}. It assumes that the DAMA count rate $s_i$ in
a time bin of width $\Delta t_i$ in an energy range $\Delta E$ for a
detector of mass $M$ is related to the muon flux $I_{\mu,i}$ during
that time by,
\begin{eqnarray}
\label{eqn:BlumModel}
s_i = \frac{y N_{\mu,i}}{M \Delta E \epsilon_i \Delta t_i},
\end{eqnarray}
where $N_{\mu,i}$ is Poisson distributed with mean,
\begin{eqnarray}
\langle N_{\mu,i} \rangle = A_{\rm eff} I_{\mu,i} \epsilon_i \Delta t_i.
\end{eqnarray}
Here $y$ is the number of signal counts/muon (yield), $A_{\rm eff}$ is
an effective area in which muons are collected, and $
\epsilon_i\approx 60\%$ is the duty cycle during time bin $t_i$. As
was argued in~\cite{Blum:2011jf}, direct hits of muons in the detector
would require $y\approx 50$ and would lead to a spread $\sigma_i/s_i$
in events which is about a factor of five larger than what is actually
observed. Since this is in clear conflict with the data, the other
possibility that we are going to consider is the secondary effect
muons can have due to the spallation reactions on nuclei and the
neutron flux they induce. In this case muons can be collected in a
larger area, say, $A_{\rm eff} \approx 10~{\rm m^2}$, which requires
an order one yield $y\approx 2$ only. Using the latter numbers we
generate $10^4$ realizations of DAMA data based on this model.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=\columnwidth]{DAMA_realizations_phase.pdf}
\end{center}
\caption{The population distribution of the phase for the model by
Blum, Eq.~(\ref{eqn:BlumModel}) compared with the best fit to data
obtained by the DAMA collaboration, $t_0 = 136$~days, shown as the
red vertical line. }
\label{fig:t0distro}
\end{figure}
We first attempt to make contact with~\cite{Blum:2011jf} by plotting
in Fig.~\ref{fig:t0distro} the distribution of $t_0$ obtained from a
$\chi^2$-fit to a sinusoidal function with floating period, phase, and
amplitude. There is little resemblance with the equivalent Fig.~3
presented in~\cite{Blum:2011jf}. The latter shows a very broad
distribution with substantial support between $90\lesssim t_0
/\mathrm{days}\lesssim 250$ and a peak at $t_0\approx
100\,\mathrm{days}$. We find from our Fig.~\ref{fig:t0distro} that
$t_0$ is normally distributed with a mean of $\langle t_0 \rangle =
187\,\mathrm{days}$ and $\sigma=27\,\mathrm{days}$. We believe that
the disagreement with Ref.~\cite{Blum:2011jf} is due to a different
choice of the time-origin. As we discussed in section~\ref{sec:phase}
once we allow the period to float, the phase $t_0$ loses its absolute
meaning and the allowed region in $t_0$ becomes sensitive to the
origin of time. By shifting the origin from Jan 1$^{\mathrm{st}}$ 2003
to the one DAMA uses when quoting their data, Jan 1$^{\mathrm{st}}$
1995, we find that the distribution in $t_0$ indeed broadens
significantly similar to the one found in~\cite{Blum:2011jf}. Though
it may certainly be better to choose the time origin in 2003, the
discrepancy signals a more serious conflict: the distribution in $t_0$
does not provide us with a robust test statistic when assessing the
compatibility of the mock data with the observed phenomena.
\begin{figure}[t]
\begin{center}
\includegraphics[width=\columnwidth]{FisherZ.pdf}
\end{center}
\caption{The population distribution of the Fisher Z-transform for the model by Blum, Eq.~(\ref{eqn:BlumModel}) as compared to the sample value of $Z = 0.47$. Based on the population distribution, the model can be excluded with a confidence exceeding $99\%$. }
\label{fig:FisherZ}
\end{figure}
A better way to assess the (in)compatibility between the muon flux and
DAMA is to use the generated set of realizations and evaluate the
degree of correlation itself. This way we remain completely coordinate
independent in our statements and we can attempt to exclude the model
by showing that it implies a level of correlation much higher than
what is actually implied by the data. This is precisely what we now
labour to show. In general, if the model allows an exact calculation
of the population correlation coefficient $r_0$ then the Fisher Z
transform,
\begin{eqnarray}
Z = \tfrac{1}{2} \log\left(\frac{1+r}{1-r} \right)
\end{eqnarray}
is approximately normally distributed with mean and standard deviation
given by,
\begin{eqnarray}
\bar{Z} = \tfrac{1}{2} \log\left(\frac{1+r_0}{1-r_0} \right)\quad\quad\quad \sigma_Z = \frac{1}{\sqrt{n-3}}.
\end{eqnarray}
We will not use these results below since the model we consider does
not easily allow for an analytic evaluation of the population
correlation coefficient $r_0$. Instead we will utilize the numerical
realizations of the model and calculate the distribution
explicitly. As it turns out, however, the numerical results nicely
agree with the theoretical expectations for the shape of the
distribution.
Figure~\ref{fig:FisherZ} shows the results of the computation of the
Fisher Z-transform for each of these realizations. The distribution is
compared to the sample's Z-transform $Z = 0.47$ obtained from
(\ref{eq:r}). The numerical realizations reveal a much larger expected
correlation than what is observed in the data. Therefore the model is
excluded with a confidence greater than $99\%$.
\subsection{CoGeNT}
\label{sec:cogent}
We now move on to investigate the hypothesis that the reported CoGeNT
signal is caused by cosmic muons. In this respect, it is important to
note that the apparent modulation fraction of the CoGeNT signal is
most prominent in the high-energy bin $1.6-3.0\,\mathrm{keV_{ee}}$
with a value in excess of 10\%~\cite{Aalseth:2011wp,Fox:2011px}.
Seasonal muon-flux modulations of that order have indeed been reported
by IceCube~\cite{Tilav:2010hj}. However, the (somewhat) milder
climatic conditions at the Minnesota Soudan Mine location exhibit a
variation of at most $4\%$~\cite{Adamson:2009zf}. Therefore, unless
the scaling of the signal is---for some unknown reason---non-linear in
incident muon flux or unless the performance of the detector was not
stable throughout the data taking period, it is not possible to
generate, say, a 16\% CoGeNT modulation from a lesser modulated
sourcing process. This statement is independent of the potential
presence of further background. Even though this argument makes a muon
explanation of CoGeNT rather unlikely, we can still proceed and look
for a quantitative answer based on the temporal structure alone.
The main complication associated with such an analysis is the fact
that although measurements of the underground muon flux are available
from the MINOS experiment~\cite{Adamson:2009zf}, the data has no
temporal overlap with CoGeNT. We circumvent this issue by relying on
nearby atmospheric temperature data rather than on measurements of the
muon flux itself. As is well known, the underground muon flux is
tightly correlated with the (stratospheric) temperature. We therefore
choose to directly evaluate Eq.~(\ref{eq:pearsonr}) between the
effective atmospheric temperature parameter $T_{\mathrm{eff}}$ and the
background subtracted CoGeNT data. We refer the reader to
Appendix~\ref{sec:cosmic-muon-flux} for further details on this
procedure.
A correlation analysis is particularly appropriate in this case since
CoGeNT's relatively short data taking period (458 days) naturally
limits the significance of any statements about the annual periodicity
of the signal as evinced in Fig.~\ref{fig:CoGeNTPeriodPhase}. However,
the correlation analysis we employ requires binning the data, a
procedure that is complicated by the need to subtract the cosmogenic
background which is responsible for the majority of the event rate in
the $0.9-1.6~{\rm keV_{ee}}$ region of the CoGeNT data. One may worry that such
subtractions introduce a bias depenedent on the choice of binning or
energy range. Moreover, the unmodulated component of the subtracted
CoGeNT data exhibits a rise towards low energies that is not shared by
the modulated part, which if anything seems larger at higher
energies. Thus, it is not even clear which part of the energy spectrum
could be a result of the muon flux. In light of these complications we
consider various time- and energy binning in the hope of thoroughly
covering the different possibilities.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=\columnwidth]{TS_Corr_tbin.pdf}
\end{center}
\caption{Coefficent of correlation $r$ between $T_{\mathrm{eff}}$ and
CoGeNT data as a function of bin size in live days. The various
fluctuating lines correspond to different binnings in energy. None
of the chosen energy regions show any significant degree of
systematic, positive correlation.}
\label{fig:rcog}
\end{figure}
Figure~\ref{fig:rcog} shows the correlation coefficient as a function
of bin size in live days for various energy ranges. If any (positive)
correlation was present, which would corroborate the hypothesis that
muons are responsible for the observed signal, one would expect
significant degree of correlation independent of the bin size. With
the coefficient of correlation fluctuating near zero, the opposite is
observed. The step-like red lines delineate the (two-sided) 90\% CL
for rejecting the null-hypothesis. The data is therefore perfectly
consistent with the null hypothesis of no correlation. For larger bin
sizes $r$ begins to fluctuate more strongly. Although it is easier to
reach some degree of correlation with fewer bins it correspondingly
becomes harder to reach a given level of significance which is
indicated by the opening of the red lines.
\section{Biannual Modulations}
\label{sec:biannual}
As discussed throughout this work, one of the hallmarks of dark matter
direct detection is the annual modulation in the recoil rate, which is
the result of the Earth's motion with and against the WIMP wind
\cite{Drukier:1986tm, Freese}. However, as we show in this section,
one also expects higher harmonics (e.g. a period of half a year) to be
present at some level and those may prove useful as additional
confidence builders if a signal is detected.
\subsection{Harmonic Analysis}
The differential recoil rate, $dR/dE_{_R}$ is a function of the Earth's
velocity relative to the rest frame of the dark matter halo, which
using the notation of Ref.~\cite{Savage:2009mk} is
approximately given by, \begin{eqnarray}
\label{eqn:vobs}
\mathbf{v}_{\rm obs} = \mathbf{v}_{\odot} + V_{\oplus}\left(\mathbf{\hat{\varepsilon}}_1 \cos \omega\left( t-t_1\right) + \mathbf{\hat{\varepsilon}}_2 \sin \omega\left(t-t_1 \right)\right)\quad \quad
\end{eqnarray}
Here $\mathbf{v}_{\odot}$ is the velocity of the sun relative to the halo, and $\mathbf{\hat{\varepsilon}}_1$ ($\mathbf{\hat{\varepsilon}}_2$) is the velocity unit vector of the Earth's motion at $t_1 = $ March 21$^\text{st}$ ($t_1+\text{year}/4$).
Due to Earth's orbit, the second term proportional to the Earth's relative velocity to the sun, $V_{\oplus} \sim 30 \text{\; km/s}$, modulates with frequency $\omega = \frac{2\pi}{1 \text{\; year}}.$ If the dark matter velocity distribution is isotropic, then the dark matter speed distribution in the earth's frame, $f(v)$, depends solely upon the magnitude of $\mathbf{v}_{\rm obs}$. To a good approximation the magnitude of the velocity is given by
\begin{eqnarray}
|\mathbf{v}_{\rm obs}| \approx |\mathbf{v}_{\odot}| + \frac{1}{2} V_{\oplus} \cos{\omega(t-t_0)}.
\label{eq:vobsmod}
\end{eqnarray}
Thus, dark matter scattering rate is a periodic function with a
fundamental period of 1 year, but like any periodic function it may
contain higher harmonic modulations of any frequency $\omega_n= n\,
\omega$. We therefore expect modulations in the scattering rate with
periods of $1/2, 1/3, \ldots$ a year corresponding to biannual,
triannual and higher frequency modulations.
In addition, there are harmonic corrections to $|\mathbf{v}_{\rm
obs}|$ itself. They arise predominantly from corrections
to~(\ref{eqn:vobs}) due to smaller effects such as the ellipticity of
the earth's motion around the sun. Once these are taken into account,
they result in observable phase shifts between the different
harmonics.
We can learn more about the higher harmonics, by expanding in the parameter $\epsilon_v = V_{\oplus}/2v_{\odot} \approx 0.06$.
The time dependence of scattering arises solely due to the velocity dependence through
\begin{eqnarray}
\nonumber \frac{dR}{dE_R}\propto &&\int^\infty_{v_{\rm min}} \frac{f(v)}{v}dv\\
\label{eq:dRtime}
&\approx& \sum_{n=0,1,\ldots} \tilde{c}_n(v_{\rm min}) \left[\epsilon_v \cos{\omega(t-t_0)}\right]^n \\ \nonumber
&=&
\sum_{n=0,1,\ldots} c_n(v_{\rm min}) \cos{\left[n\omega(t-t_0)\right]}
\end{eqnarray}
Here we have used the fact that trigonometric identities allow powers of
$\cos{[\omega(t-t_0)]}$ to be re-expressed as sums of $\cos{\left[ n
\omega(t-t_0)\right]}.$ In Eq.~(\ref{eq:dRtime}), $v_{\rm min}$ is the
minimum dark matter speed in the lab which can deposit a recoil energy
$E_R$,
\begin{eqnarray}
\label{eqn:vmin}
v_{\rm min} = \frac{1}{\sqrt{2m_N E_R}} \left(\frac{m_N E_R}{\mu_{N\chi}} +\delta \right)
\end{eqnarray}
for a nuclear target of mass $m_N$ and dark matter-nuclei reduced mass $\mu_{N\chi} = m_N m_\chi/(m_N+m_\chi)$; in~(\ref{eqn:vmin}) we have included the possibility of inelastic scattering with a splitting $\delta$ in the dark matter sector \cite{iDM,MiDM}.
The computation of the harmonic coefficients requires a choice of velocity distribution, and in what follows we consider the distribution proposed in~\cite{Lisanti:2010qx},
\begin{eqnarray}
f_k(v) \propto \left[ \exp\left(\frac{v_{\rm esc}^2 - v^2}{kv_0^2} \right) - 1\right]^k\Theta\left(v_{\rm esc} - v\right)
\end{eqnarray}
Here, $\Theta(x)$ is the Heaviside step function, and we consider the dispersion $v_0 = 220 \text{\; km/s}$~\cite{Bovy:2009dr,McMillan:2009yr,Reid:2009nj} and the escape velocity $v_{esc} = 600 \text{\; km/s}$~\cite{Smith:2006ym}.
In Fig.~\ref{fig:amplitudes} (top), we illustrate the behavior of the
harmonic coefficients, $c_n$, for the $k$=1.5 velocity distribution.
In order to plot the amplitudes on a logarithmic scale, we have taken
the absolute values of $c_n$. The troughs in the plots indicate when
$c_n$ is changing sign. At high enough minimum velocity, $v_{\rm min}$,
the harmonic coefficients are all positive. Interestingly, there
are~$n$ sign changes for~$c_n$, which can be understood from the behavior of
the velocity distribution~$f(v)$.
In the approximation of Eq.~(\ref{eq:dRtime}), the phases of all the
higher harmonics are the same as the annual modulation. However, as
mentioned above, there are higher order corrections to the magnitude
of the observer's velocity, Eq.~(\ref{eq:vobsmod}), that allow the
phases of the higher harmonics to deviate from the phase of the annual
modulation. Indeed, using an accurate parameterization of the earth's
velocity~\cite{Lewin:1995rx}, we find appreciable temporal shifts of
the biannual and triannual mode with respect to the annual one. This
is shown in the middle panel of~Fig.~\ref{fig:amplitudes}. The abrupt
changes seen in the figure indicate sign changes in the harmonic
coefficients $c_n$, which can be thought of as phase shifts
by~$1~\rm{yr}/2n$. The ambiguities in the phase shifts are fixed by
choosing the values which are closest to the annual one. These phase
shifts serve as an additional signature which can be used to
discriminate a potentially positive signal from other sources.
In the bottom plot of Fig.~\ref{fig:amplitudes}, we plot the
modulation amplitude ratio $c_n/c_1$ for the biannual and triannual
modulation divided by the annual modulation. There are two regions
where the higher harmonics are as important as the annual modulation.
The first region occurs near 200 km/s around the zero in the annual
modulation coefficient~$c_1$. Here, the biannual amplitude~$c_2/c_0$
is at the per mille level which is small but potentially observable
with a large amount of exposure. The second region where higher
harmonics are important is at high velocities, above the escape
velocity cutoff. At these high velocities, there are substantially
less dark matter particles in the winter than in the summer. Thus,
the scattering rate can vanish entirely in winter, with a chopped
cosine behavior, which requires more harmonics to fully describe the
approach to zero. Although this would be obvious in a experiment with
no background, the presence of background can mask these higher
harmonics by eye, so it is useful to look for them.
We have also looked at the behavior of these modulation amplitudes for
the Standard Halo Model and also for a $k=3.5$ distribution
\cite{Lisanti:2010qx}. The behavior at low $v_{\rm min}$ is very similar,
since that region depends mostly on the velocity dispersion and not on
the behavior near the tail. The high velocity behavior enhances
(suppresses) the higher harmonics for sharper (shallower) escape
velocity cutoffs as seen in the $k=3.5$ distribution (Standard Halo
Model).
A detailed analysis of the detectability of these higher harmonics and
their ability to be faked by backgrounds is beyond the scope of this
paper. Here, we restrict ourselves to a few remarks. Any background
that modulates annually, with a small modulation amplitude will,
through the same Taylor expansion argument, also have higher harmonics
(although not necessarily all with similar phase). What is peculiar
about dark matter is that the higher harmonic amplitudes are enhanced
at high $v_{\rm min}$ due to the escape velocity physics. This could be
mimicked by background only if there was some effect during winter
which severely suppressed the background appearing as a nuclear
scattering event.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=0.96\columnwidth]{TS_biannualPHASE_revised.pdf}
\end{center}
\caption{The velocity dependence of dark matter scattering as a
function of $v_{\rm min}$ for the $k$=1.5 velocity distribution proposed
in \cite{Lisanti:2010qx}, with dispersion $v_0 = 220 \text{\; km/s}$
and escape velocity $v_{esc} = 600 \text{\; km/s}$. The top plot
shows the $|c_{n}|$ values for n=$0,1,2,3$ (top to bottom) where
$c_n$ is in units of (km/s)$^{-1}$, the middle plot is the phase
shift of the higher harmonics relative to the annual modulation,
while the bottom plot shows the modulation amplitudes
$|c_{n}/c_{1}|$ for n=$2,3$ (top to bottom). The horizontal lines in
the bottom panel show the 90\% C.L. upper limits on a bi- and
triannual signal in the DAMA data.}
\label{fig:amplitudes}
\end{figure}
On the theory side, models with inelastic dark matter scattering lead
to enhanced higher harmonic amplitudes as compared with standard
elastic dark matter. A splitting $\delta \sim 100$~keV leads to values
of $v_{\rm min}$ which can be near the cutoff (see Eq.~\ref{eqn:vmin}).
In fact, we find that the biannual amplitude can be as large as 30\%
of the annual modulation amplitude for the 2-4 keV$_{ee}$ bins of
DAMA. Looking at Fig.~\ref{fig:DAMA-LIBRA_null}, we see that this is
consistent with the amount of biannual power seen in these bins, since
the power ratio between biannual and annual modulation is $(c_2/c_1)^2
\lesssim 0.1.$ With a large increase in exposure, by a factor of
$O(10-100)$, DAMA would have enough statistics to begin testing this
and perhaps even see evidence for a nonzero biannual amplitude as we
show quantitatively in the next subsection.
As a final remark, we note that it is possible to predict the energy
bin where the annual modulation should be suppressed, due to its
change of sign, which would be an interesting bin to look for biannual
modulation. For heavy dark matter that scatters elastically, the
dependence on the dark matter mass drops out of the reduced mass, so
we can use Eq.~\ref{eqn:vmin}, to show that the energy bin $E_R = 90
\text{\; keV} (m_N/100 \text{\; GeV})$ is where the annual modulation
should be suppressed. Such high energies might be hard to get substantial dark matter rates though, since the nuclear form factors tend to suppress such high energy recoils. The situation gets better for dark matter much lighter than the nucleus, as the energy bin is $E_R = 0.9 \text{\; keV} (m_\chi/10 \text{\; GeV})^2 (m_N/100 \text{\; GeV})^{-1}.$ This requires knowledge of the dark matter mass to predict, but experimentally it might be verified that the sign of the annual modulation amplitude changes sign in a certain bin, thus it would be an interesting followup to see if there is a biannual mode in that bin. In fact, the DAMA annual modulation amplitude is smaller in the first bin $2-2.5 ~{\rm keV_{ee}}$ bin, so it would be interesting to see if there is any biannual modulation there.
\subsection{Testing the Signal Hypothesis}
In the presence of a signal, the probability distribution function of the power $P$ in a particular frequency of the Lomb-Scargle power spectrum changes to~\cite{1975ApJS...29..285G,1982ApJ...263..835S},
\begin{eqnarray}
p(P; P_s) = I_0(2\sqrt{P ~P_s}) \exp\left(-P-P_s\right).
\end{eqnarray}
This formula assumes a normalized signal with power $P_s$ that is measured in units of the variance of the noise.
Given an observation of power $P_{\rm obs}$ we can reject a signal at a level of $1-\alpha$ by requiring that,
\begin{eqnarray}
{\rm Pr}(P>P_{\rm obs}) = 1 - \int_0^{P_{\rm obs}} p(P;P_s) dP = 1-\alpha
\end{eqnarray}
So, for example, from Figure \ref{fig:DAMA-LIBRA_null}, the DAMA/LIBRA
data in the $2-4~{\rm keV_{ee}}$ energy range has a power of $P_{\rm obs}
=0.57$ (1.8) in the biannual (triannual) mode. This implies a
frequentist 90\% CL upper limit of $P_s^{\rm biannual} <1.9\, (4.4)$
on the power of the biannual (triannual) mode of a putative
signal. Since DAMA claims detection of an annually modulating signal
we can find the ratio of the higher harmonic power to the annual
power. The power observed in the annual mode of the same energy bin is
$11.0$. Therefore the fractional power in the biannual or triannual mode must be
lower than,
\begin{align}
P_s^{\rm biannual} &< 0.17 P_s^{\rm annual} , \nonumber \\
P_s^{\rm triannual} &< 0.4 P_s^{\rm annual}\quad \text{ at 90\% CL} .
\end{align}
Since the power is directly related to the square of the harmonic
coefficient, this limit implies a corresponding limit on the harmonic
coefficients, $c_2/c_1 < 0.42$ and $c_3/c_1 < 0.63$. These limits are
shown in the bottom panel of Fig.~\ref{fig:amplitudes}.
A similar analysis for the muon flux gives a 90\% CL upper limit on
the biannual mode of $P_s^{\rm biannual} < 6.3$, see Figure
\ref{fig:LVD_null}. Since the power observed in the annual cycle is
$P_s^{\rm annual} = 312$ this implies a limit on the harmonic
coefficients of $c_2/c_1 < 0.17$.
\section{Conclusions}
\label{sec:conclusions}
This paper was devoted to a detailed time-series analysis of Dark Matter direct detection data as well as related datasets.
The main findings are as follows: there is no evidence
to support the idea that atmospheric muons are responsible for the
signal that the DAMA collaboration observes; the annual modulations
observed in the CoGeNT data are statistically significant only in the
higher part of the low energy spectrum ($1.6-3.0~{\rm keV_{ee}}$); no
correlation of the latter with the expected muon flux is observed;
biannual modulations may serve as an additional handle on any putative
dark matter signal.
With regards to the muon hypothesis discussed in refs.~\cite{ralston,
nygren, Blum:2011jf} whereby atmospheric muons reaching the DAMA
detector are responsible for the signal observed, our analysis
indicates three significant difficulties with this idea:
\begin{enumerate}
\item The power spectrum of the LVD data on the muon flux indicates
considerably more power at periods longer than one year compared
with the DAMA/LIBRA data. This remains true even after we follow the
undesirable procedure of the DAMA collaboration whereby the
residuals are computed by subtracting the mean of the data on a
yearly basis. This procedure is undesirable since it masks power at
long periods. We therefore urge the DAMA collaboration to release
the unsubtracted data to allow for a full comparison of the power
spectrum. Similar conclusions have been reached
in~\cite{Blum:2011jf}.
\item When fitting the two datasets to a sinusoidal function with
variable amplitude, phase, and period the most likely regions for
the two fits do not overlap. On this point we are in disagreement
with ref.~\cite{Blum:2011jf} which finds that the phase and period
of the muon dataset can be in agreement with the DAMA signal. We
believe that the source of this discrepancy is the choice of time
origin which becomes important once the period is allowed to float.
\item A correlation analysis which is independent of any choice of
functional form or fit to the datasets excludes the simple, yet very
general model presented by ref.~\cite{Blum:2011jf} for connecting
the muon flux with the DAMA signal. It indicates that such a model
would imply a much stronger degree of correlation than what is in
fact observed in the data.
Thereby, this analysis also puts to question the
hypothesis~\cite{ralston,nygren} itself that the variation of the
muon flux has a casual connection with the one seen by DAMA.
\end{enumerate}
Aside from these three difficulties, which depend only on the temporal
structure of the two signals, there also seems to be some difficulty
reconciling the amplitude of the muon modulations with that of the
DAMA signal. On the face of it, the amplitude of the modulations
observed by LVD seems to agree well with the $\sim2\%$ modulations as
originally quoted by the DAMA collaboration. However, the full
single-hit energy spectrum as measured in
DAMA-LIBRA~\cite{Bernabei:2008yi}, seems to suffer from $^{40}$K
contamination in the signal region. Moreover, it was recently pointed
out in ref.~\cite{Kudryavtsev:2010zza} that the radioactive background
rates in the DAMA experiment are likely larger than what was estimated
by DAMA. If true, that would imply a smaller unmodulated amplitude of
the putative signal and hence a larger modulation fraction. Thus, any
model attempting to explain the DAMA signal as a result of the muon
flux will require some sort of nonlinearity in the signal yield.
Given all of the above, we find the muon hypothesis in its current
form rather unpersuasive. Any further investigations into this
possibility should address the aforementioned difficulties.
With regards to the CoGeNT data our analysis concentrated on three
different energy regions, $0.5-0.9~{\rm keV_{ee}}$, $0.9-1.6~{\rm keV_{ee}}$, and
$1.6-3.0~{\rm keV_{ee}}$. We find that the presence of annual modulations can
only be firmly established in the $1.6-3.0~{\rm keV_{ee}}$ region. The other
two regions require a somewhat more careful treatment of the data to
remove the contribution from the cosmogenic activity which contributes
to the power spectrum at long periods due to the long-lived isotopes
involved. Nevertheless, none of the procedures we have followed for
subtracting the cosmogenic contributions resulted in a statistically
significant yearly modulation signal in the subtracted data. The
phase analysis of the modulations shows them to peak around the
middle of April, which is somewhat earlier than expected from DM recoils, but the uncertainty on this number is still rather large. Indeed, we find that the confidence intervals in the full two dimensional phase-period space are much too large for any significant statement to be made about the modulations observed.
As in the case of DAMA we find no significant correlation between the
CoGeNT data set and the yearly modulation of the muon flux. In the absence of published
measurements of the latter for the time-period of the CoGeNT run we
used the atmospheric effective temperature inferred from
climate data instead. The one-to-one relation between the variation in
temperature and muon-flux allows us again to indirectly disfavor the
muon hypothesis as it applies to the modulations seen by CoGeNT.
One can also wonder if muon induced backgrounds instead have a delayed signal, would that allow them to explain these dark matter anomalies? As shown in Appendix~\ref{sec:activation}, production of long-lived radioactive isotopes does not result in any improvement. The largest allowed temporal shift of the signal peak is to the beginning of October. To complicate matters, the modulation amplitude of such a signal is always smaller than the muon amplitude. These limitations make it difficult to explain DAMA or CoGeNT with such muon backgrounds.
Finally, we discussed a new diagnostic for dark matter direct detection experiments. Since the expected recoil rate of dark matter
against matter in the lab is a general periodic function with a
fundamental period of one year, it also contains biannual modulations
as well as higher harmonics. The relative amplitude of the velocity
modulation is roughly $V_{\oplus}/2v_{\odot} \approx 0.06$, which is
rather small. The higher harmonics in the recoil rate have similar
phase to the annual modulation, but are generally suppressed.
However, we presented a few cases where they might be observed given
sufficient exposure. At the very least, the biannual mode can be
looked for and limits on its power can be obtained in a
straightforward fashion.
\begin{acknowledgments}
We acknowledge helpful discussions with L.~Dai, M.~Kamionkowski, D.~Morrissey and
M.~Pospelov. We thank J.~Collar for assistance in interpreting the
CoGeNT data.
\end{acknowledgments}
|
\section{Introduction}
Equilibrium statistical mechanics describes the connection of a few
macroscopic intensive quantities, e.g. pressure and temperature, with the
microscopic properties of the many particles constituting the
system~\cite{mcquarrie}. Although there is no equivalent formalism for systems
driven out of thermal equilibrium studies are often motivated by the quest for
simple governing principles such as an effective
temperature~\cite{cugl97a,cugl11}. Of particular interest are small
mesoscopic systems such as colloidal particles, nanoparticles in solution, or
biological systems, all of which are dominated by fluctuations.
For systems only slightly perturbed from equilibrium -- into what is called
the \emph{linear response regime} -- the fluctuation-dissipation theorem (FDT)
relates the response with equilibrium correlations through the
temperature~\cite{kubo}. This unique temperature corresponds to what we would
measure with a thermometer, and is independent of the observables entering the
fluctuation-dissipation theorem. Examples for these observables are the
velocity related to the diffusion coefficient, or stress fluctuations
determining the viscosity. The practical importance of the FDT both for
experiments and simulations thus stems from the possibility to extract
transport properties from stationary fluctuations. In a more recent
application the FDT has been used to predict the self-assembly of model
systems~\cite{jack07,klot11}.
Even for systems driven into a non-equilibrium steady state (NESS) far beyond
the linear response regime we can still define a linear response
function~\cite{agar72,hang82,cris03,marc08}. For the broad class of driven
systems with Markovian dynamics and embedded in a fluid at well defined
temperature it has been demonstrated recently that then the FDT is most
convincingly interpreted in terms of an \emph{excess}
correlation~\cite{spec06,blic07,spec09,baie09,pros09,seif10,spec10}. The
single temperature that enters these generalized FDTs is that of the fluid,
and no approximations are involved. Nevertheless, for the purpose of a simple
description we might still be interested in defining an approximate
temperature through the fluctuation-dissipation ratio (FDR), i.e., the ratio
of correlation to response function. This strategy has originally been
proposed in the context of aging mean-field spin
systems~\cite{cugl97a,cugl11}, and subsequently been applied to many different
systems~\cite{ono02,haya04,haxt07}. In particular, in a sheared colloidal
suspension or fluid the Einstein relation between the self-diffusion
coefficient of a tagged particle and its mobility is broken and can be used to
define an effective temperature~\cite{bert02,szam04,krug09,land10,zhan11}.
\begin{figure}[b!]
\centering
\includegraphics[width=.9\linewidth]{models.png}
\caption{(Color online) Systems studied: (A)~Toy model with a single
particle bound to the origin and (B)~tagged particle in a dense
suspension. Both systems are driven through linear shear flow.}
\label{fig:models}
\end{figure}
In this paper we study a many-body system governed by underdamped stochastic
dynamics and driven into a NESS through linear shear flow. Such a system could
model a fluid with every particle coupled to a stochastic thermostat,
colloidal or nano-suspensions, or dusty plasmas~\cite{shuk01}. We follow a
tagged particle, i.e., a randomly chosen particle out of many identical
interacting particles. Motivated by the physical picture of an effective
confinement in dense systems we also consider a single trapped colloidal
particle in shear flow~\cite{zieh09} as a toy model, see
Fig.~\ref{fig:models}. We discuss the FDT in a ``hybrid'' form: we relate
response and correlations through the kinetic temperature in the spirit of an
effective temperature, but with an additive correction term still present. For
the trapped particle expressions can be obtained analytically. For the tagged
particle we derive a similar FDT exploiting a time-scale separation due to the
effective confinement. In both cases we show that the correction term indeed
becomes negligible for strong confinement.
\section{Fluctuation-dissipation ratio}
We study the response of a single particle with mass $M$ moving in a viscous
liquid at temperature $T$. After applying a small external force $\vec f$
directly to the particle its mean velocity evolves as
\begin{equation}
\label{eq:v:R}
\mean{\vec v(t)} = \IInt{t'}{0}{t} \mat R(t-t')\vec f(t') + \mathcal O(f^2),
\end{equation}
where $\mat R=(R_{ij})$ is the response matrix with components
\begin{equation}
\label{eq:R}
R_{ij}(t-t') \equiv \left.\fd{\mean{v_i(t)}}{f_j(t')}\right|_{\vec f=0}
\end{equation}
with $t\geqslant t'$ due to causality. The brackets $\mean{\cdots}$ denote the
thermal average in the perturbed system. Throughout this paper we employ
dimensionless units and measure length in units of the particle diameter $a$
and energy in units of $k_\text{B} T$. Time is measured in units of
$\tau_0\equiv a^2/D_0$, which quantifies the time it takes for a free particle
with diffusion coefficient $D_0$ to diffuse a distance equal to its
diameter. The reduced mass $m\equiv(MD_0/k_\text{B} T)/\tau_0$ is the ratio of
the momentum relaxation time to the diffusive time scale $\tau_0$.
We describe the stochastic particle motion through the coupled equations
$\dot\x=\vec v$ and
\begin{equation}
\label{eq:lang}
m\dot{\vec v} = -\nabla U + \vec f - [\vec v-\vec u(\x)] + \boldsymbol\xi,
\end{equation}
where $\x$ and $\vec v$ are the particle position and velocity,
respectively. Besides the conservative forces arising from the potential $U$
we can perturb the particle by a direct force $\vec f$. The noise $\boldsymbol\xi$
modeling the interactions of the particle with solvent molecules has zero mean
and correlations
\begin{equation}
\label{eq:nois}
\mean{\boldsymbol\xi(t)\boldsymbol\xi^\text{T}(t')} = 2\;\mathbf 1\delta(t-t').
\end{equation}
The external shear flow enters through the term $\vec u(\x)=\shr y\vec e_x$,
i.e., the flow points in $x$-direction and increases its amplitude linearly
with the $y$-coordinate. Here, $\shr$ is the strain rate which in our units is
equal to the P\'eclet number.
Eq.~\eqref{eq:R} measures the linear response of the system. If the
unperturbed system ($\vec f=0$) is in thermal equilibrium the FDT
\begin{equation}
\label{eq:fdt}
\mat R(t-t') = \mat C(t-t') \equiv \mean{\vec v(t)\vec v^\text{T}(t')}_0
\end{equation}
relates this response to the velocity auto correlation function (VACF) $\mat
C(t)$, where the subscript indicates that these correlations are to be
measured in the unperturbed system.
For a computationally convenient representation of the response function
Eq.~\eqref{eq:R} valid both in and out of equilibrium consider the path weight
of the noise
\begin{equation}
P[\boldsymbol\xi(t)] \sim \exp\left\{ -\frac{1}{4}\Int{t} [\boldsymbol\xi(t)]^2 \right\}.
\end{equation}
The stochastic velocity of the particle is a result of previous collisions
with solvent molecules, $\vec v(t)=\vec v[t;\boldsymbol\xi(\tau)]$. Since a force
perturbation is equivalent to perturbing the noise we can write
\begin{equation}
R_{ij}(t-t') =
\int[\boldsymbol\xi(\tau)] \fd{v_i[t;\boldsymbol\xi(\tau)]}{\xi_j(t')}P[\boldsymbol\xi(\tau)]
\end{equation}
for the response. A functional integration by parts then leads
to~\cite{spec06,cala05}
\begin{equation}
\label{eq:R:nois}
\mat R(t-t') = \frac{1}{2}\mean{\vec v(t)\boldsymbol\xi^\text{T}(t')}_0.
\end{equation}
Hence, even if the system is driven into a NESS the response can be measured
through a steady-state correlation function (see also
Refs.~\cite{chat04,bert07}). However, the FDT in the form Eq.~\eqref{eq:fdt}
no longer holds and the dimensionless fluctuation-dissipation ratio (FDR) is
defined as
\begin{equation}
\label{eq:fdr}
X_i(t) \equiv \frac{C_{ii}(t)}{R_{ii}(t)}
\end{equation}
for the diagonal components.
To calculate $X_i(0)$ we perform a short-time expansion of Eq.~\eqref{eq:v:R}
leading to $\mean{\vec v}\approx\Delta t\mat R(0)\vec f$. From the equation of
motion~\eqref{eq:lang} we obtain $m\mean{\dot{\vec v}}_0\approx m\mean{\vec
v}/\Delta t=\vec f$. We have exploited that the average force on the
particle right before the perturbation vanishes, $\mean{\vec{\nabla}U}_0=0$,
which is quite obvious for isotropic systems but due to the inversion symmetry
about the origin it holds also in the presence of simple shear flow. Hence, in
our dimensionless units we obtain $R_{ij}(0)=\delta_{ij}/m$ and finally
\begin{equation}
\label{eq:Tk}
X_i(0) = m\mean{v_i^2}_0 \equiv \theta_i.
\end{equation}
The right hand side is the \emph{kinetic temperature} as measured through the
velocity fluctuations. In the following we study for two systems whether, and
under which conditions, Eq.~\eqref{eq:Tk} extends to times $t>0$, i.e.,
whether $X_i(t)\approx\theta_i$.
\section{Trapped particle}
The toy model we investigate first is a single particle trapped in the
harmonic potential
\begin{equation}
\label{eq:harm}
U(r) = \frac{1}{2}k r^2
\end{equation}
with strength $k$, where $\x$ is the displacement from the origin and
$r=|\x|$. Due to the linearity of the restoring force the $z$-component in
Eq.~\eqref{eq:lang} decouples and remains in equilibrium,
$X_z(t)=\theta_z=1$. Therefore, in this section we only consider the motion in
the $xy$-plane.
\subsection{Analytical results}
Due to the quadratic potential Eq.~\eqref{eq:harm} the equations
of motion comprise a linear system of first order differential
equations. Hence, we can solve it for the velocity
\begin{equation}
\label{eq:harm:v}
\vec v(t) = \mat G^{vr}(t)\x_0 + \mat G^{vv}(t)\vec v_0
+ \frac{1}{m}\IInt{t'}{0}{t}\mat G^{vv}(t-t')\boldsymbol\xi(t')
\end{equation}
given the initial displacement $\x_0$ and initial velocity $\vec v_0$. The
explicit expressions for the Green's functions $\mat G(t)$ are given in the
appendix~\ref{sec:green}. Both the VACF and the response function are easily
calculated from the solution Eq.~\eqref{eq:harm:v}. For the VACF we find
\begin{equation}
\label{eq:harm:C}
\mat C(t) = \mean{\vec{v}(t)\vec{v}_0^\text{T}}_0 =
\mat G^{vr}(t)\mean{\vec{r}_0\vec{v}_0^\text{T}}_0 + \mat G^{vv}(t)
\mean{\vec{v}_0\vec{v}_0^\text{T}}_0,
\end{equation}
while the response function is trivially related to the Green's function
through
\begin{equation}
\label{eq:harm:R}
\mat R(t-t') = \frac{1}{2}\mean{\vec v(t)\boldsymbol\xi^\text{T}(t')}_0
= \frac{1}{m}\mat G^{vv}(t-t')
\end{equation}
using the noise correlations Eq.~\eqref{eq:nois}. From the steady state
distribution Eq.~\eqref{eq:sim:psi} we obtain the moments
\begin{gather}
\mean{\x_0\vec v_0^\text{T}}_0 = \frac{1}{2k}\left(
\begin{array}{cc}
0 & -\shr \\ \shr & 0
\end{array}\right),
\\
\mean{\vec v_0\vec v_0^\text{T}}_0 = \frac{1}{m}\mathbf 1
+ \frac{1}{2k}\left(
\begin{array}{cc}
\shr^2 & 0 \\ 0 & 0
\end{array}\right).
\end{gather}
The kinetic temperatures Eq.~\eqref{eq:Tk} perpendicular to the shear flow are
$\theta_y=\theta_z=1$, whereas
\begin{equation}
\label{eq:harm:Tk}
\theta_x = 1 + \alpha_x\shr^2 \geqslant 1, \qquad \alpha_x \equiv \frac{m}{2k}.
\end{equation}
As a consequence of the linear forces and the symmetry of the shear flow the
excess compared to equilibrium is proportional to $\shr^2$.
\begin{figure}[t]
\centering
\includegraphics{trapped.pdf}
\caption{(Color online) Single particle moving in a harmonic trap: (A-C)
scaled velocity auto-correlation function $C_{xx}(t)/\theta_x$ and response
function $R_{xx}(t)$ \textit{vs.} time $t$ for strain rate $\shr=2$ and
different masses $m$ and trap strengths $k$. (B,C)~For large $k$, the
correction in Eq.~\eqref{eq:harm:corr} becomes negligible and both curves
lie on top of each other. (D)~The magnitude of the correction term
$\beta_x$ as a function of $k$ for two different masses. (E)~Sketch of the
different regimes of the FDT for $m=0.6$. The dashed lines $\shr^2$ and
$1/m$ limit the region where the FDR is approximatly time-independent,
$X_x\approx\theta_x$. Below the solid line $m\shr^2$ the kinetic temperature
is much larger than unity. While for the chosen $m$ there is a gap (shaded
area), with increasing $m$ both regimes can be realized in the vicinity of
the solid line.}
\label{fig:harm}
\end{figure}
\subsection{The FDT}
Using the explicit expressions for the moments we see that with
$G^{vr}_{yx}(t)=0$ the correlation function for the $y$-component reads
$C_{yy}(t)=G^{vv}_{xx}\mean{v_y^2}_0$ and therefore $X_y(t)=\theta_y=1$ at all
times. On the other hand, in the direction of the shear flow we obtain
\begin{equation}
\label{eq:harm:corr}
C_{xx}(t) = \theta_xR_{xx}(t) + \mean{yv_x}_0 G^{vr}_{xy}(t),
\end{equation}
i.e., the velocity correlations are expressed through the response times the
kinetic temperature plus a correction term. Separating the dependence on
strain rate the correction term can be rewritten as
\begin{equation}
\label{eq:beta}
\mean{yv_x}_0 G^{vr}_{xy}(t) = \shr^2\beta_x I_x(t),
\end{equation}
where $\max|I_x(t)|=1$ and $\beta_x$ captures the magnitude of the correction
term. In Fig.~\ref{fig:harm}, we plot scaled response and correlation
functions for three representative values of the parameters $m$ and $k$. As
demonstrated in Fig.~\ref{fig:harm}D increasing the trap strength $k$ strongly
decreases $\beta_x$. The behavior of the FDT is sketched in
Fig.~\ref{fig:harm}E, where we compare the magnitude of the correction term to
the kinetic temperature. From Eq.~\eqref{eq:harm:Tk} we see that for $k\ll
m\shr^2$ we have $\theta_x\gg 1$ (see also Fig.~\ref{fig:harm}A). For the
relevant case $k\gtrsim 1/m$ we find from the explicit expressions that for
the FDR to become approximately time-independent $k\gg\max\{\shr^2,m\shr^2\}$
must hold (Fig.~\ref{fig:harm}C). While for small masses $m<1$ there is a gap
(this case is sketched in Fig.~\ref{fig:harm}E), for sufficiently large
$m\gtrsim1$ these two regimes come close. In Fig.~\ref{fig:harm}B, we
demonstrate that there is indeed a regime of intermediate trap strength $k\sim
m\shr^2$ with an increased effective temperature where nevertheless
$X_x(t)\approx\theta_x$ holds to a very good degree.
\section{Tagged particle in a suspension}
The main system we now study is a suspension composed of $N$ particles in
which we tag and follow a single particle, say $k=1$, with position $\x_1$ and
velocity $\vec v_1$. The particles interact through a pair potential $u(r)$
with total potential energy $U=\sum_{i<j}u(|\x_i-\x_j|)$. Advection through
the shear flow leads to the well-known Taylor
dispersion~\cite{tayl53}. Instead of the absolute velocity $\vec v_1$ it will
be more convenient to use the relative velocity
\begin{equation}
\vec v \equiv \vec v_1 - \vec u(\x_1)
\end{equation}
with respect to the flow as the observable entering the response $\mat R(t)$
and correlation function $\mat C(t)$.
\begin{figure*}[t]
\centering
\includegraphics{susp.pdf}
\caption{(Color online) Tagged particle with $m=1$ in a sheared suspension
at volume fractions $\phi=0.1$ (top row) and $\phi=0.4$ (bottom row):
(A)~Pair distribution function in the $xy$-plane for strain rate
$\shr=1$. (B)~Scaled velocity auto-correlation functions $C_{xx}(t)/\theta_x$
and response functions $R_{xx}(t)$ for different strain rates: from bottom
to top $\shr=0,0.2,0.4,0.6,0.8,1$. For visibility curves are
shifted. (C)~Off-diagonal response functions $R_{xy}(t)$ (solid lines) and
$R_{yx}(t)$ (dashed lines) for the same strain rates and offsets. (D)~The
magnitude of the correction term $|\theta_{xy}R^\ast_{ij}|$, i.e., the
product of Eq.~\eqref{eq:Tk:off} with the maximum $R^\ast_{ij}$ of the
off-diagonal response function, as a function of strain rate. The dashed
lines show the quadratic fits.}
\label{fig:susp}
\end{figure*}
\subsection{Time-scale separation}
In the following we assume a time scale separation between the motion of the
potential energy minimum (or inherent state position~\cite{stil84})
$\x_\text{c}$, and the vibrational motion of the tagged particle around
$\x_\text{c}$. The physical picture is that particles vibrate in a ``cage'' of
surrounding particles and that local reorganization takes much longer than the
vibrational motion. Linearizing the force exerted by neighboring particles on
the tagged particle leads to
\begin{equation}
-\nabla_1 U \approx -\mat k[\x_1(t)-\x_\text{c}(t)], \qquad
k_{ij} \equiv \left.\pd{^2U}{r_i\partial r_j}\right|_{\x=\x_\text{c}}.
\end{equation}
We solve the resulting equations of motion leading to the same formal result
Eq.~\eqref{eq:harm:C} for the correlation function and Eq.~\eqref{eq:harm:R}
the response function. In principle the Green's function can be calculated but
we will not need its explicit form here. To see that the first term in
Eq.~\eqref{eq:harm:C} vanishes consider the projected probability
\begin{equation}
\bar\psi_\text{s}(\x_1,\vec v) = \Int{\x_2\cdots\mathrm{d}\x_N}\Int{\vec v_2\cdots\mathrm{d}\vec
v_N}\psi_\text{s}(\{\x_i\})
\end{equation}
of the tagged particle, where $\psi_\text{s}$ is the full stationary distribution. For a
homogeneous, translationally invariant system $\bar\psi_\text{s}$ cannot depend on the
position $\x_1$ and therefore $\mean{\x_1\vec v^\text{T}}_0\sim\mean{\vec v}_0=0$
vanishes. Hence, we obtain
\begin{equation}
\label{eq:susp:corr}
C_{ii}(t) = \theta_i R_{ii}(t) + m\sum_{j\neq i}\mean{v_iv_j}_0R_{ij}(t)
\end{equation}
as our central result. The correction term differs from
Eq.~\eqref{eq:harm:corr} and now couples to the off-diagonal elements of the
response function instead of the off-diagonal component of the Green's
function connecting velocity and position.
\subsection{Langevin dynamics simulations}
We perform Langevin dynamics simulations to study Eq.~\eqref{eq:susp:corr} for
a specific system. The $N=1728$ particles are enclosed in a cubic simulation
box with edge length $L$. The particles interact through the Yukawa (or
screened Coulomb) pair potential
\begin{equation}
\label{eq:yuk}
u(r) =
\begin{cases}
\varepsilon\frac{e^{-\kappa(r-1)}}{r} & (r\geqslant 1) \\
\infty & (r<1),
\end{cases}
\end{equation}
where $\varepsilon$ is the interaction energy at contact and $\kappa^{-1}$ the
screening length determined by the composition of the surrounding solvent. We
choose $\varepsilon=8.0$ and $\kappa=5.0$ in order to obtain a broad range of
densities for which the liquid phase is stable~\cite{azha00}. For the shear
flow we employ Lees-Edwards boundary conditions, which enforce a linear
velocity profile in the suspension~\cite{allen}. We integrate the equations of
motion by a stochastic velocity Verlet algorithm~\cite{frenkel} with a time
step $5\times10^{-4}$. Since the hard-core repulsion cannot be implemented in
the interaction potential we employ a simple algorithm that detects collisions
and computes the appropriate positions and velocities after the impact
according to momentum and energy conservation (see Refs.~\cite{stra99,foss00}
and references therein). The NESSs are prepared by initializing the particle
positions on a regular lattice at low density. Then we equilibrate the system
and slowly increase the density by decreasing the box size. After this
equilibrium system is constructed we slowly ramp up the strain rate until the
target value $\shr$ is reached. We simulate another $1000$ time steps to relax
the system into the steady state. This procedure is repeated separately for
independent runs.
We study suspensions at different volume fractions $\phi\equiv\pi N/(6L^3)$
and strain rates $\shr$. To determine the response and correlation functions
we simulate the motion of the system in the NESS and record the velocity
trajectories of $200$ randomly chosen particles in four independent runs. From
this data we can easily evaluate the VACF. To compute the response to a small
force we again employ Eq.~\eqref{eq:R:nois}, which is still valid for the many
particle system. This enables us to obtain the response function from steady
state trajectories without the need to explicitly perturb the
system. Therefore, in addition to the velocity, we also record the stochastic
forces acting on the tagged particles, which are directly accessible in a
numerical simulation.
\subsection{Numerical results}
The pair distribution function quantifies the probability to find another
particle at a displacement $\x$. It is distorted from its isotropic
equilibrium shape, see Fig.~\ref{fig:susp}A. This function visualizes the
average environment of the tagged particle. Under the influence of the shear
flow there is a higher number of particles in the compressional zones and a
depletion in the extensional zones~\cite{foss00}. Moreover, the peak
corresponding to the first nearest-neighbor shell becomes more pronounced.
\begin{figure}[b!]
\centering
\includegraphics{low.pdf}
\caption{(Color online) Dilute suspension at $\phi=0.01$ and strain rate
$\shr=1$ (left) and $\shr=2$ (right). In comparison to Fig.~\ref{fig:susp}
a deviation between response $R_{xx}(t)$ and velocity autocorrelation
function $C_{xx}(t)$ is observed.}
\label{fig:low}
\end{figure}
In Fig.~\ref{fig:susp}B, the response functions $R_{xx}(t)$ together with the
scaled VACF $C_{xx}(t)/\theta_x$ is plotted for volume fractions $\phi=0.1$ and
$\phi=0.4$, and for a range of strain rates. Even for times $t>0$ these
functions coincide, which implies that the additive correction term in
Eq.~\eqref{eq:susp:corr} vanishes. The off-diagonal response functions are
plotted in Fig.~\ref{fig:susp}C. For comparison response and correlation
functions are plotted in Fig.~\ref{fig:low} for a dilute suspension at volume
fraction $\phi=0.01$. Here we observe a clear difference between correlations
and response. To understand the observed behavior one should bear in mind that
we consider the velocities relative to the local flow. At low densities
collisions are rare and the particles adapt smoothly to the flow of the
solvent. Therefore, the largest deviations from the flow profile arise when a
particle diffuses in $y$-direction entering a region of faster or slower flow
in $x$-direction, to which it needs to adapt. For higher densities collisions
become much more frequent. These collisions prevent the particles from
adapting to the solvent flow. Additionally they distribute the momentum,
transfered to the particles by the shear flow, more or less randomly in the
three spacial directions. The result is that under shear flow the diagonal
velocity moments grow with increasing density, and become more and more
similar in size.
The two components $\mean{v_xv_z}_0\simeq0$ and $\mean{v_yv_z}_0\simeq0$ of
the velocity correlation matrix are very small (see also
Fig.~\ref{fig:mass}). Defining the off-diagonal ``temperature''
\begin{equation}
\label{eq:Tk:off}
\theta_{xy} \equiv m\mean{v_xv_y}_0
\end{equation}
we see that the dominant contribution to the correction term in
Eq.~\eqref{eq:susp:corr} is $\theta_{xy}R_{xy}(t)$ for the $x$ component, and
$\theta_{xy}R_{yx}(t)$ for the $y$ component. In analogy to the trapped particle
[Eq.~\eqref{eq:beta}] we separate the strain rate dependence of the correction
terms,
\begin{equation}
\theta_{xy}R_{xy}(t) \approx \shr^2\beta_x I_x(t), \quad
\theta_{xy}R_{yx}(t) \approx \shr^2\beta_y I_y(t),
\end{equation}
with again $\max|I_i(t)|=1$ and coefficients $\beta_i$. In
Fig.~\ref{fig:susp}D, we plot $|\theta_{xy}R^\ast_{ij}|\approx\shr^2\beta_i$,
where $R^\ast_{ij}$ is the maximal value of the off-diagonal component of the
response matrix. For $\phi=0.1$ we observe the predicted quadratic dependence
on the strain rate $\shr$. For $\phi=0.4$ the quadratic predicition holds for
$\shr\leqslant0.8$, while for larger strain rates higher order terms in $\shr$
become important.
\begin{figure}[t]
\centering
\includegraphics{teff.pdf}
\caption{(Color online) The kinetic temperatures $\theta_i$ (symbols) and their
quadratic fits (lines) plotted \textit{vs}. strain rate $\shr$ for
$m=1$. Shown are the three directions for volume fractions (A)~$\phi=0.1$
and (B)~$\phi=0.4$. In (C) the coefficients $\alpha_i$ and
$\beta_x\times10^2$ are plotted as function of the volume fraction
$\phi$. (D)~Relative magnitude of the correction term
$|\theta_{xy}R^\ast_{xy}|/\theta_x\times10^3$ as function of strain rate and
density, cf. Fig.~\ref{fig:harm}E. Colors are linearly interpolated.}
\label{fig:teff}
\end{figure}
In Fig.~\ref{fig:teff}, we plot the kinetic temperatures as a function of
strain rate. While for $\phi=0.1$ a clear difference between motion parallel
to the flow ($\theta_x$) and motion perpendicular to the flow
($\theta_y\simeq\theta_z$) can be seen, this distinction is diminished at higher
densities. Moreover, all kinetic temperatures can be well fitted by the
quadratic function Eq.~\eqref{eq:harm:Tk} with coefficients $\alpha_i$ for the
three directions. The increase of velocity fluctuations can be explained by
forced collisions due to the flow gradient, an effect that is more pronouned
at higher densities and higher strain rates.
In Fig.~\ref{fig:teff}C, the coefficients $\alpha_i$ and $\beta_x$ are shown for
the different densities. The coefficient $\beta_x$ decreases for larger
densities but then turns up again at $\phi=0.4$. The reason for this
non-monotonic behavior is that there are two effects determining the shape of
the $xy$-component of the response function, see Fig.~\ref{fig:resp}. One
dominates for lower, one for higher densities. At low densities forcing the
particle upwards in $y$-direction moves it into a region of faster flow. While
the velocity relaxes the relative velocity with respect to the shear flow is
negative. At higher densities this effect is weaker because of more frequent
particle collisions, and thus large excursions in $y$-direction are
rare. Hence, the velocity differences due to the motion are smaller and so is
the response caused by this effect. In addition another effect of the
collisions becomes more significant. Pulling the particle upwards in
$y$-directions makes collisions with particles from the left more likely than
with particles from the right. This leads to an average acceleration to the
right which counteracts the first effect. In the intermediate density regime
these two effects almost cancel and lead to a small $\beta_x$.
\begin{figure}[t]
\centering
\includegraphics[width=.9\linewidth]{response.png}
\caption{(Color online) Sketch of the time evolution after a force
perturbation in $y$-direction (arrow) determining the shape of the
$R_{xy}(t)$ response function: (A)~At low densities the tagged particle is
slower than the surrounding flow field due to inertia. (B)~At high
densities collisions with neighboring particles are more likely, pushing
the tagged particle in the direction of the flow.}
\label{fig:resp}
\end{figure}
Finally, in Fig.~\ref{fig:teff}D the magnitude of the correction term with
respect to the response function, $|\theta_{xy}R^\ast_{xy}|/\theta_x$, is plotted,
see also Fig.~\ref{fig:harm}E. Note that over the whole parameter range
studied here this value is lower than $0.02$, i.e., $X_x\approx\theta_x$ holds to
a very good degree. However, we see that for large strain rates and low
density this ratio grows by an order of magnitude. Due to the effects
described above for the highest density the range of strain rates for which
the correction term is negligible shrinks again.
\subsection{Overdamped limit}
Our results have been derived for systems with underdamped stochastic
dynamics. Of greater practical importance in colloidal suspensions is the
overdamped limit corresponding to neglecting inertia, $m\ra0$. In
Fig.~\ref{fig:mass}, the dependence of the coefficients on the reduced mass $m$
is plotted. As expected for smaller $m$ the kinetic temperatures approach
unity, $\theta_i\approx1$. The additive correction term is no longer
negligible as $\beta_x$ and $\beta_y$ grow. We thus recover the
fluctuation-dissipation theorem as described in the introduction, in which the
bath temperature enters and the equilibrium form of the FDT is completed by an
excess correlation function.
Inserting the overdamped equation of motion for the tagged particle into
Eq.~\eqref{eq:R:nois} one arrives at
\begin{equation}
\label{eq:ov}
C_{ij}(t) = 2R_{ij}(t) + \mean{F^{(1)}_i(t)F^{(1)}_j(0)}_0
\end{equation}
for $t>0$ and components $i,j=y,z$ perpendicular to the shear
flow~\cite{land10}. Here, $\vec F^{(1)}=-\nabla_1U$ is the force on the tagged
particle exerted by its neighboring particles. The time-integrated version of
Eq.~\eqref{eq:ov} has been obtained previously within mode-coupling
calculations~\cite{szam04,krug09,krug10}. For systems in which the force-force
correlations can be neglected Eq.~\eqref{eq:ov} predicts a universal FDR
$X_i=2$. It has been argued that such an approximation might be justified in
dense suspensions close to the glass transition under sufficiently large
shear.
\begin{figure}[t]
\centering
\includegraphics{mass.pdf}
\caption{(Color online) The coefficients $\beta_i$ (solid lines, left axis)
and the kinetic temperature $\theta_x$ (dashed line, right axis) as a
function of the reduced mass $m$ for volume fraction $\phi=0.4$ and strain
rate $\shr=2$.}
\label{fig:mass}
\end{figure}
\section{Summary}
We have studied the relation between velocity autocorrelation and response
function of a tagged particle moving in a suspension that is driven into a
non-equilibrium steady state through simple shear flow. Under the assumption
of a time-scale separation between vibrational motion and local reorganization
the tagged particle effectively behaves like a trapped particle. The diagonal
components of the tagged particle's velocity autocorrelation function are then
given by
\begin{equation}
C_{ii}(t;\shr) \approx \theta_i(\shr)R_{ii}(t;\shr) + \shr^2\beta_iI_i(t;\shr)
\end{equation}
with expansion of the kinetic temperature
\begin{equation}
\theta_i(\shr) \approx 1+\alpha_i\shr^2.
\end{equation}
Here, $R_{ii}(t;\shr)$ are the response functions Eq.~\eqref{eq:R},
$I_i(t;\shr)$ are functions of order unity, and $\alpha_i$ and $\beta_i$ are
coefficients independent of the strain rate $\shr$. While these expressions
are exact for the trapped single particle our numerical results show that they
hold approximately to a very good degree for a tagged particle moving in an
interacting colloidal suspension. One might of course anticipate the quadratic
dependence on strain rate from symmetry arguments close to equilibrium. We
have shown here that these expressions follow from a time-scale separation
caused by an effective confinement.
The effect of the shear flow is to break symmetry and to couple the particle
velocity to earlier perturbations perpendicular to its velocity. This leads to
an additive correction term that grows with $\shr^2$. Since $\beta_i\ll1$ this
correction is negligible up to dimensionless strain rates
$\shr\sim(\max\{\beta_i\})^{-1/2}$, which can be far from
equilibrium. Specifically, here we have studied strain rates in the range
$\shr\leqslant1$ and found excellent agreement between correlation and
response functions, see Fig.~\ref{fig:susp}. However, already at
$\shr\gtrsim2$ the kinetic temperatures start to divert from the quadratic
law, indicating the importance of higher order terms. Increasing the density
the tagged particle interacts more strongly with its surrounding
particles. The kinetic temperature increases due to more frequent collisions
with neighboring particles in conjunction with transport due to the flow. This
is in contrast to the trapped particle where tightening the trap reduces
fluctuations and therefore the kinetic temperature approaches the bath
temperature.
An intriguing perspective is to apply our results to supercooled (or
supersaturated) conditions. Since dynamics slows down dramatically and the
time-scale separation between vibrations and long-lived particle displacements
becomes even more pronounced we expect that our results extend into the
supercooled regime. While we have employed stochastic dynamics one might
speculate that our results also hold in systems governed by deterministic
dynamics such as the SLLOD equations of motion~\cite{tuck97} as employed in
Ref.~\cite{bert02}. Future work will also address the influence of
hydrodynamic interactions and flow generated through boundaries.
\acknowledgments
We acknowledge financial support by the DFG through project SE~1119/3. TS
acknowledges funding through Alexander von Humboldt foundation and, during
early stages of the project, by the Director, Office of Science, Office of
Basic Energy Sciences, Materials Sciences and Engineering Division and
Chemical Sciences, Geosciences, and Biosciences Division of the
U.S. Department of Energy under Contract No.~DE-AC02-05CH11231.
|
\section{Introduction}
The standard model (SM) of elementary particles is a successful
theory without any contradiction to the observations up to now.
However, it contains many free parameters, most of
which come from Yukawa couplings.
The eigenvalues and eigenvectors
of the Yukawa coupling matrices determine the mass ratios and mixings
between generations, respectively, and their observed values are quite
hierarchical.
Models beyond the SM should explain such hierarchical structures
of quarks and leptons.
Extra dimensions provide a simple way to realize
the hierarchical flavor structures, \ie,
a wave function localization of matter fields
in extra dimensions~\cite{Arkani-Schmaltz}.
Actually, the most promising
candidate for the unified theory of the SM and the gravity,
\ie, the superstring theory, predicts the existence of extra dimensions.
They can also explain other problems of the SM, such as
the gauge hierarchy problem~\cite{RS},
a candidate for dark matter~\cite{Servant:2002aq}, and so on.
Supersymmetry (SUSY) is an interesting extension of the SM.
It softens the divergences in quantum field theories
and then protects the electroweak scale against large radiative
corrections.
The three SM gauge couplings are unified in the minimal supersymmetric
standard model (MSSM) at $M_{\rm GUT}\equiv 2\times 10^{16}$~GeV,
which suggests the grand unified theory.
It also has a candidate for dark matter if the R-parity forbids decays
of the lightest SUSY particle.
Besides, the existence of SUSY is predicted by the superstring theory.
\ignore{
Among various extensions of the SM, supersymmetric extensions are quite
interesting because, in addition to the fact that superstring theory
predicts supersymmetry (SUSY), it softens the divergences in quantum
field theories and then protects the electroweak (EW) scale against
the radiative corrections. Furthermore, the superpartners of SM particles
would contain dark matter which are required by cosmological observations,
and the three gauge coupling constants of the minimal supersymmetric
standard model (MSSM) are unified at the so-called grand unification
theory (GUT) scale.
}
From the above reasons, we consider an extension of the SM by
introducing SUSY and extra dimensions.
The minimal setup for such an extension is a five-dimensional (5D)
supergravity (SUGRA) compactified on an orbifold~$S^1/Z_2$.
The chiral structure of the SM can be realized
by the orbifold $Z_2$ projection, which preserves $N=1$ SUSY
in a four-dimensional (4D) sense.
The local SUSY (\ie, SUGRA) is required in order to
avoid the existence of a massless Goldstino which contradicts to many
observations, and to discuss the moduli stabilization.
The off-shell formulation of 5D SUGRA~\cite{Kugo-Ohashi,Kugo:2002js}
provides the most general way to construct 5D SUGRA models.
There is, moreover, a systematic way to obtain 4D effective theories of
such models, which we call the off-shell dimensional
reduction~\cite{Abe:2007} based on the $N=1$ superfield description
of 5D SUGRA~\cite{Paccetti Correia:2004ri,Abe:2004}.
This method can be applied to general 5D SUGRA models.
For example, we analyzed some class of models
by this method~\cite{Abe:2007zv}
that include SUSY extension of the Randall-Sundrum model~\cite{SUSY_RS}
and the 5D heterotic M theory~\cite{5D_Mtheory} as special limits
of parameters.
The wave function localization of matter fields (hypermultiplets)
is realized by controlling bulk mass parameters for the matter fields
in 5D models.
In 5D SUGRA on $S^1/Z_2$, the bulk mass parameters are obtained
as $U(1)$ charges (times 5D Planck mass) under $Z_2$-odd vector
multiplets including the graviphoton multiplet.
One of the minimal models with the realistic flavor structure
was constructed with a single $Z_2$-odd vector multiplet.
However, it suffers from the SUSY flavor problem as well as
tachyonic squarks and/or sleptons~\cite{Abe:2008ka}.
In our previous paper~\cite{Abe:2008an},
we constructed models with two $Z_2$-odd vector multiplets, which
induce a modulus chiral multiplet other than the radion multiplet
in the 4D effective theory, and showed that
there is an important contribution of the multiple moduli multiplets
to the effective K\"{a}hler potential that may solve problems
mentioned above.
In this paper, we extend the previous work to more generic set-up,
which has an arbitrary number of $Z_2$-odd vector multiplets
(\ie, moduli multiplets) and a nontrivial warping along
the extra dimension.
We perform a comprehensive and systematic analysis to understand
the SUSY flavor structure of such generic 5D SUGRA models, and
provide a phenomenological analysis based on an illustrative model.
This paper is organized as follows.
In Sec.~\ref{review_5DSUGRA}, we set up our model
with a brief review of the off-shell formulation of 5D SUGRA.
In Sec.~\ref{4Deffective}, we derive 4D effective theory
of our 5D model and study its properties.
In Sec.~\ref{specific_model},
we perform a phenomenological analysis based on an illustrative model.
Sec.~\ref{summary} is devoted to a summary.
In Appendix~\ref{derive:L_eff}, some details of the derivation
of the effective action are shown.
In Appendix~\ref{comment:4-fermi}, we provide a comment on some peculiar
structure of the 4D effective theory to 5D SUGRA.
In Appendix~\ref{explicit_forms}, explicit expressions of some quantities
in the model in Sec.~\ref{specific_model} are collected.
\section{Set-up and brief review of 5D SUGRA} \label{review_5DSUGRA}
\subsection{Brief picture of set-up} \label{setup}
The set-up we consider in this article is as follows.
\begin{itemize}
\item The fifth dimension is compactified on the orbifold~$S^1/Z_2$,
and the background 5D metric is the warped metric.
In contrast to the original warped model~\cite{RS},
the warp factor is supposed to be $\cO(10^2)$ and explains
the small hierarchy between the Planck scale~$M_{\rm Pl}=2.4\times 10^{18}$~GeV
and the GUT scale~$M_{\rm GUT}=2\times 10^{16}$~GeV.
\item The compactification scale is around $M_{\rm GUT}$,
and below this scale, 4D effective theory becomes MSSM.
\item All the standard model fields are identified with
zero-modes of the 5D bulk fields.
\item The hierarchical flavor structure of the SM fermions
is realized by the quasi-localization of the wave functions
for the zero-modes~\cite{Arkani-Schmaltz}.
\item SUSY is broken at some scale below $M_{\rm GUT}$, and
the main source of SUSY breaking is the $F$-term of
a single chiral superfield~$X$ in 4D effective theory.
\end{itemize}
When we study a model with extra dimensions,
it is indispensable to stabilize the size of the compactified internal space
to a finite value.
Especially in SUSY models, such stabilization mechanisms affect
the sfermion mass spectrum in the MSSM sector.
In order to take into account the stabilization of the extra dimension,
we have to work in the context of SUGRA.
Although the above type of set-up has been studied in many papers,
most of them do not consider the full SUGRA effects
or only consider a {\it limited} case from the viewpoint of SUGRA.
In our previous work~\cite{Abe:2008an}, we pointed out
a possibility to solve the SUSY flavor problem
thanks to some peculiar terms
in the K\"{a}hler potential of 4D effective theory,
in a case that the theory has two moduli
multiplets (we will provide the definition of the moduli multiplet
in the next subsection).
In this article, we extend our previous work
to the case of an arbitrary number of the moduli multiplets
and a non-trivial warp factor, and discuss some phenomenological
aspects.
Before specifying the model, we start with {\it general} 5D SUGRA,
derive 4D effective theory in a systematic way,
and analyze the flavor structure of the soft SUSY breaking terms
in the effective theory.
After that, we will construct an illustrative model
in Sec~\ref{specific_model} to realize the above set-up.
For our purpose, the superconformal gravity formulation,
which is an off-shell description of 5D SUGRA, is useful.
So we briefly review this formulation
and explain the structure of 5D SUGRA
in the rest of this section.
\subsection{Superconformal multiplets in 5D SUGRA}
The background 5D metric with 4D Poincar\'{e} invariance is parametrized as
\be
ds^2 = e^{2\sgm(y)}\eta_{\mu\nu}dx^\mu dx^\nu-dy^2,
\ee
where $\mu,\nu=0,1,2,3$, $\eta_{\mu\nu}=\diag(1,-1,-1,-1)$,
and $e^{\sgm(y)}$ is a warp factor, which is a function of only $y$
and determined by the dynamics.
We take the fundamental region of the orbifold as $0\leq y\leq L$.
Since we choose the coordinate~$y$ such that $\vev{e_y^{\;\;4}}=1$,
the constant~$L$ denotes the size of the extra dimension.
Our formalism is based on the superconformal formulation
developed in Ref.~\cite{Kugo-Ohashi,Kugo:2002js}.
5D superconformal multiplets relevant to our study are summarized
in Table~\ref{sc_multiplets}.
\begin{table}[t]
\begin{center}
\begin{tabular}{l|c|c}
\rule[-2mm]{0mm}{7mm} 5D superconformal multiplet &
$N=1$ decomposition & $Z_2$-parity \\ \hline\hline
\rule[-2mm]{0mm}{7mm} Weyl multiplet (gravity) &
$\bdm{E_W}=(E_W,L^\alp_E,V_E)$ & $(+,-,+)$ \\ \hline
\rule[-2mm]{0mm}{7mm} Vector multiplet (moduli) &
$\bdm{V}^{I'}=(V^{I'},\Sgm^{I'})$ & $(-,+)$
\\ \hline
\rule[-2mm]{0mm}{7mm} Vector multiplet (gauge)&
$\bdm{V}^{I''}=(V^{I''},\Sgm^{I''})$ & $(+,-)$ \\ \hline
\rule[-2mm]{0mm}{7mm} Hypermultiplet (compensator) &
$\bdm{H}^{a=1}=(\Phi^1,\Phi^2)$ & $(-,+)$ \\ \hline
\rule[-2mm]{0mm}{7mm} Hypermultiplet (matter) &
$\bdm{H}^{a\geq 2}=(\Phi^{2a-1},\Phi^{2a})$ & $(-,+)$ \\ \hline
\end{tabular}
\end{center}
\caption{Relevant 5D superconformal multiplets. Each multiplet is
decomposed into $N=1$ multiplets. }
\label{sc_multiplets}
\end{table}
Each multiplet can be decomposed into $N=1$ multiplets~\cite{Kugo:2002js}.
The signs in the last column of Table~\ref{sc_multiplets}
denote the orbifold $Z_2$ parities of the $N=1$ multiplets
in the next column.
We assume that each $N=1$ multiplet has the same $Z_2$ parity
at both boundaries~$y=0,L$, for simplicity.\footnote{
In general, vector and hypermultiplets can have different
$Z_2$ parities at different boundaries,
but only $Z_2$-even fields at both boundaries have zero-modes
which are relevant to 4D effective theory.
All the other fields are decoupled and should be integrated out.
Thus the treatment of multiplets with different $Z_2$ parities
is the same as that of multiplets which are $Z_2$-odd at both boundaries. }
\begin{description}
\item[Weyl multiplet~$\bdm{E_W}$]
This corresponds to the gravitational multiplet,
and is decomposed into the $N=1$ Weyl multiplet~$E_W$,
a complex general multiplet~$L_E^\alp$ ($\alp$: spinor index),
and a real general multiplet~$V_E$.
Among them, $E_W$ and $V_E$ are $Z_2$-even,
and includes the 4D parts of the vierbein~$e_\mu^{\;\;\underline{\nu}}$
and the extra-dimensional component~$e_y^{\;\;4}$, respectively.
The $Z_2$-odd multiplet~$L_E^\alp$ includes
the ``off-diagonal'' parts~$e_\mu^{\;\;4}$ and $e_y^{\;\;\underline{\mu}}$.
The latter is irrelevant to the following discussion, and is neglected.
When the loop corrections are taken into account, however,
the contribution from $L_E^\alp$ has to be included.
\item[Vector multiplet~$\bdm{V}^I$]
This is decomposed into
$N=1$ vector and chiral multiplets~$V^I$ and $\Sgm^I$,
which have opposite $Z_2$-parities.
The vector multiplets are divided into two classes
according to their $Z_2$ parities.
One is a class of the gauge multiplets, which are denoted as $\bdm{V}^{I''}$.
In this class, $V^{I''}$ are $Z_2$-even and have zero-modes
that are identified with the gauge multiplets in 4D effective theory.
In the other classes ($\bdm{V}^{I'}$),
the 4D vector components have no zero-modes.
Instead, the chiral multiplets~$\Sgm^{I'}$ have zero-modes.
They include the scalar fields and their potential is flat
at the classical level.
Thus we refer to $\Sgm^{I'}$ (or $\bdm{V}^{I'}$) as the moduli multiplets
in this article.\footnote{
These moduli fields are actually identified with the shape moduli
of the compactified space for a 5D effective
theory of the heterotic M-theory compactified
on the Calabi-Yau manifold~\cite{5D_Mtheory}, for example. }
At least one vector multiplet belongs to the latter category.
In the pure SUGRA, the vector component of such a multiplet
is identified with the graviphoton.\footnote{
In this article, the terminology~``graviphoton'' represents
the vector field in the gravitational multiplet of
the {\it on-shell} formulation.
It should be distinguished from the off-diagonal components of
the 5D metric, which are included in $L_E^\alp$ in the current formulation. }
All the other components are auxiliary fields that are eliminated
by the superconformal gauge-fixing.
Thus, when there are $n_V$ moduli multiplets in the off-shell action,
only $(n_V-1)$ degrees of freedom are physical.
(See Sec.~\ref{sc_GF}.)
\item[Hypermultiplet~$\bdm{H}^a$]
This is decomposed into
two chiral multiplets~$\Phi^{2a-1}$ and $\Phi^{2a}$,
which have opposite $Z_2$-parities.
We can always choose their $Z_2$-parities
as listed in Table~\ref{sc_multiplets}
by using $SU(2)_U$, which is an automorphism of the superconformal algebra.
The hypermultiplets are also divided into two classes.
One is the compensator multiplets~$a=1,2,\cdots,n_C$ and the other is
the physical matter multiplets~$a=n_C+1,\cdots,n_C+n_H$.
The former is an auxiliary degree of freedom and eliminated
by the superconformal gauge-fixing.
In contrast to 4D SUGRA, both types of hypermultiplets have the same
quantum numbers of the superconformal symmetries in 5D SUGRA.
They are discriminated only by signs of their kinetic terms in the action.
Thus, in principle, it is possible to introduce
an arbitrary number of the compensator multiplets in the theory.
In this article, we consider the single compensator case ($n_C=1$)
for simplicity.\footnote{
The number of the compensator multiplets determines the target
manifold of the hyperscalars.
For instance, it is
$USp(2,2n_H)/USp(2)\times USp(2n_H)$ for $n_C=1$,
and $SU(2,n_H)/SU(2)\times U(n_H)$ for $n_C=2$. }
\end{description}
\subsection{$\bdm{N=1}$ description of 5D action}
For our purpose, it is convenient to describe the 5D action
in terms of the $N=1$ multiplets~\cite{Paccetti Correia:2004ri,Abe:2004}.
This corresponds to the extension of the result in Ref.~\cite{ArkaniHamed:2001tb}
to the local SUSY case.
We can see that $V_E$ has no kinetic term in this description.\footnote{
This does not mean that $e_y^{\;\;4}$ is an auxiliary field.
It is also contained in $\Sgm^I$, which have their own kinetic terms. }
After integrating it out, the 5D Lagrangian is expressed
as~\cite{Correia:2006pj}
\bea
\cL \eql -3e^{2\sgm}\int\dr^4\tht\;\cN^{1/3}(\cV)
\brc{d_{\hat{a}}^{\;\;\hat{b}}\bar{\Phi}_{\hat{b}}
\brkt{e^{-2g_It_IV^I}}^{\hat{a}}_{\;\;\hat{c}}\Phi^{\hat{c}}}^{2/3}
\nonumber\\
&&-e^{3\sgm}\sbk{\int\dr^2\tht\;\Phi^{\hat{a}}d_{\hat{a}}^{\;\;\hat{b}}
\rho_{\hat{b}\hat{c}}\brkt{\der_y-2ig_It_I\Sgm^I}^{\hat{c}}_{\;\;\hat{d}}
\Phi^{\hat{d}}+\hc} \nonumber\\
&&+\cL_{\rm vec}+2\sum_{y_*=0,L}\cL^{(y_*)}\dlt(y-y_*), \label{5D_action}
\eea
where $d_{\hat{a}}^{\;\;\hat{b}}=\diag(\id_{2n_C},-\id_{2n_H})$,
$\rho_{\hat{a}\hat{b}}=i\sgm_2\otimes\id_{n_C+n_H}$,
$\bar{\Phi}_{\hat{b}}\equiv (\Phi^{\hat{b}})^\dagger$,
and $\cL_{\rm vec}$ is defined as
\be
\cL_{\rm vec} \equiv \int\dr^2\tht\brc{
-\frac{\cN_{IJ}(\Sgm)}{4}\cW^I\cW^J
+\frac{\cN_{IJK}}{48}\bar{D}^2\brkt{
V^ID^\alp\der_y V^J-D^\alp V^I\der_y V^J}\cW^K_\alp}+\hc. \label{def:L_CS}
\ee
This contains the kinetic terms
for the vector multiplets~$V^I$ and the Chern-Simons terms.
The indices~$\hat{a},\hat{b}$ run over the whole $2(n_C+n_H)$ chiral
multiplets coming from the hypermultiplets.
As mentioned in the previous subsection,
we consider the case of $n_C=1$ in the following.
Here $\sgm_2$ in $\rho_{\hat{a}\hat{b}}$ acts
on each hypermultiplet~$(\Phi^{2a-1},\Phi^{2a})$.
$\cN$ is a cubic polynomial called the norm function,\footnote{
This corresponds to the prepotential in the $N=2$ global SUSY case. }
which is defined by
\be
\cN(X) \equiv C_{IJK}X^IX^JX^K.
\ee
A real constant tensor~$C_{IJK}$ is completely symmetric for the indices,
and $\cN_I(X)\equiv \der\cN/\der X^I$,
$\cN_{IJ}(X)\equiv\der^2\cN/\der X^I\der X^J$, and so on.
The superfield strength~$\cW^I_\alp\equiv -\frac{1}{4}\bar{D}^2D_\alp V^I$
and $\cV^I\equiv -\der_y V^I+\Sgm^I+\bar{\Sgm}^I$
are gauge-invariant quantities.
The generators~$t_I$ are anti-hermitian.
For a gauge multiplet of a non-abelian gauge group~$G$,
the indices~$I$, $J$ run over $\dim G$ values and $\cN_{IJ}$
are common for them.
The index~$a$ for the hypermultiplets are divided into irreducible
representations of $G$.
The fractional powers in the first line of (\ref{5D_action})
appear after integrating $V_E$ out.
The boundary Lagrangian~$\cL^{(y_*)} (y_*=0,L)$ can be introduced
independently of the bulk Lagrangian.
Note that (\ref{5D_action}) is a shorthand expression of the full SUGRA action.
We can always restore the full action by the promotion,
\be
\int\dr^4\tht\brc{\cdots} \to \frac{1}{2}\sbk{\cdots}_D, \;\;\;\;\;
\int\dr^2\tht\brc{\cdots}+\hc \to \sbk{\cdots}_F, \label{FDpromote}
\ee
where $\sbk{\cdots}_D$ and $\sbk{\cdots}_F$ denote the $D$- and $F$-term action
formulae of the $N=1$ superconformal formulation~\cite{4Doffshell},
which are compactly listed in Appendix~C of Ref.~\cite{Kugo:2002js}.\footnote{
The superfield descriptions of the $D$- and $F$-term formulae
on the ordinary superspace are provided
at the linear order in the fields belonging to $E_W$
in Ref.~\cite{Sakamura:2011df}. }
Here we omitted the spacetime integral.
This promotion restores
the dependence of the action on the components
of the $N=1$ Weyl multiplet~$E_W$,
such as the Einstein-Hilbert term and the gravitino-dependent terms.
The Weyl multiplet also contains some auxiliary fields.
After integrating them out, some terms in (\ref{5D_action}) are modified.
Practically, the kinetic terms for $V^I$ are the only such terms
that are relevant to the phenomenological discussions.
Their kinetic functions are read off as
$-\cN_{IJ}(\Sgm)/4$ from the first term in (\ref{def:L_CS}).
This will be modified after integrating out the above-mentioned
auxiliary fields, and
the correct kinetic function is obtained as $\brc{\cN a_{IJ}}(\Sgm)/2$,
where~\cite{Kugo-Ohashi}
\be
a_{IJ} \equiv -\frac{1}{2\cN}\brkt{\cN_{IJ}-\frac{\cN_I\cN_J}{\cN}}.
\label{def:a_IJ}
\ee
\subsection{Superconformal gauge-fixing in 5D} \label{sc_GF}
Here we provide some comments on the superconformal gauge-fixing in 5D SUGRA.
Since we will not impose the gauge-fixing conditions
at the 5D stage,\footnote{
The superconformal gauge-fixing will be imposed after
4D effective action is derived because it breaks the $N=1$ off-shell
structure of the action. }
the readers can skip this subsection.
Nevertheless, the comments presented here may help the readers
to understand the structure of 5D SUGRA set-up.
The Lagrangian~(\ref{5D_action}) with the promotion~(\ref{FDpromote})
is invariant (up to total derivatives)
under the superconformal symmetries.
In order to obtain the usual Poincar\'{e} SUGRA,
we have to eliminate the extra symmetries by imposing
the gauge-fixing conditions.
A conventional choice of such conditions is expressed
in our $N=1$ superfield notation as
\bea
\left.\cN\brkt{\frac{\cV}{V_E}}\right|_0
\eql \left.
d_{\hat{a}}^{\;\;\hat{b}}\bar{\Phi}_{\hat{b}}\Phi^{\hat{a}}\right|_0 = M_5^3,
\nonumber\\
\left.\cN\brkt{\frac{\cV}{V_E}}\right|_\tht \eql
\left.\cN\brkt{\frac{\cV}{V_E}}\right|_{\bar{\tht}^2\tht} =
\left.d_{\hat{a}}^{\;\;\hat{b}}\bar{\Phi}_{\hat{b}}\Phi^{\hat{a}}
\right|_\tht =
\left.\Phi^{\hat{a}}d_{\hat{a}}^{\;\;\hat{b}}\rho_{\hat{b}\hat{c}}\Phi^{\hat{c}}
\right|_\tht = 0, \;\;\; \cdots, \label{GF_cond}
\eea
where $M_5$ is the 5D Planck mass,
and the symbols~$|_0$, $|_\tht$ and $|_{\bar{\tht}^2\tht}$
denote the lowest, $\tht$- and $\bar{\tht}^2\tht$-components
in the superfields, respectively.
$V_E=e_y^{\;\;4}+\tht\psi_y^++\bar{\tht}\bar{\psi}_y^++\cdots$
is the real general multiplet
coming from the 5D Weyl multiplet (see Table~\ref{sc_multiplets}),
where $\psi_y^+$ is the $Z_2$-even 5th-component of the gravitino.
The conditions in the first line fix the dilatation,
and those in the second line fix the conformal SUSY.
They reproduce the Einstein-Hilbert
term~$\cL=-\frac{M_5^3}{2}e^{(5)}\cR^{(5)}+\cdots$,
where $e^{(5)}$ is the determinant of the f\"unfbein,
$\cR^{(5)}$ is the 5D Ricci scalar, from the $D$-term action formula.
The conditions in (\ref{GF_cond}) indicate that
there is one multiplet whose components are not physical
in each of the vector and hypermultiplet sectors.
Such a multiplet is the graviphoton multiplet in the vector multiplet sector,
and the compensator multiplet in the hypermultiplet sector.
However, the graviphoton~$B_M$ ($M=\mu,y$) itself is exceptional.
Since (\ref{GF_cond}) does not involve the vector components,
the graviphoton is always physical.
The first condition in (\ref{GF_cond}) suggests that
$\Sgm^{I'}|_0$ generically have nonvanishing VEVs.
(Since $\Sgm^{I''}$ are $Z_2$-odd, they do not have zero-modes nor VEVs.)
Specifically, the lowest components of $V^I$ and $\Sgm^I$ are
\be
V^I = \tht\sgm^\mu\bar{\tht}W_\mu^I+\cdots, \;\;\;\;\;
\Sgm^I = \frac{1}{2}\brkt{e_y^{\;\;4}M^I-iW_y^I}+\cdots,
\ee
where $M^I$ and $W^I_M$ are the real scalar and vector components
of the 5D vector multiplet~$\bdm{V}^I$,
and the 4D vector part of the graviphoton multiplet is identified as
\be
V_B \equiv \frac{\cN_I}{3\cN}(2\Re\vev{\Sgm})V^I, \label{def:V_B}
\ee
whose fermionic component vanishes by (\ref{GF_cond}).
The corresponding chiral multiplet part is defined as
\be
\cT \equiv \frac{\cN_I}{3\cN}(2\Re\vev{\Sgm})\Sgm^I. \label{def:cT}
\ee
In contrast to $V_B$, this remains physical under the gauge-fixing conditions.
In fact, (\ref{GF_cond}) and (\ref{def:V_B}) suggest that
\be
\cT = \frac{1}{2}\brkt{e_y^{\;\;4}-iB_y}+\tht\psi_y^++\cdots,
\label{comp:cT}
\ee
where $B_M=\frac{\cN_I}{3\cN}(2\Re\vev{\Sgm})W^I_M$ is the graviphoton.
We have chosen the coordinate~$y$ such that $\vev{e_y^{\;\;4}}=1$.
\ignore{
\be
\cN(2\Re\Sgm)|_0 = \brkt{M_5 e_y^{\;\;4}}^3, \;\;\;\;\;
\cN_I(2\Re\Sgm)|_0 \Sgm^I|_\tht = 3M_5^3\brkt{e_y^{\;\;4}}^2\psi_y^+.
\ee}
Eq.(\ref{comp:cT}) shows that $\cT$ no longer belong
to the vector multiplet sector
after the gauge-fixing~(\ref{GF_cond}),
but it is the ``5D radion multiplet'' belonging to the 5D gravitational sector.
Therefore, among $\Sgm^I$, one combination~$\cT$ has a different origin
from the others.
We can also see this fact explicitly from the action.
The condition~(\ref{GF_cond}) suggests that $\Phi^2$ must have
a nonzero VEV, and it plays a similar role to the chiral compensator
multiplet in 4D SUGRA.
($\Phi^1$ does not have a VEV because it is $Z_2$-odd.)
To emphasize this point, we rewrite the hypermultiplets as
$\phi\equiv\brkt{\Phi^2}^{2/3}$ and
$\hat{\Phi}^{\hat{a}}\equiv\Phi^{\hat{a}}/\Phi^2$ ($\hat{a}\neq 2$)
so that their Weyl weights are one and zero, respectively.
Then the first line of (\ref{5D_action}) is expanded around $\vev{\Sgm^I}$ as
\bea
\cL_D \eql -3\hat{\cN}^{1/3}(\vev{\cV})
\int\dr^4\tht\;\abs{\phi}^2\brc{
1+\frac{\cN_I}{\cN}(\vev{\cV})\tl{\cV}^I
+\frac{3\cN\cN_{IJ}-2\cN_I\cN_J}{18\cN^2}(\vev{\cV})
\tl{\cV}^I\tl{\cV}^J}+\cdots \nonumber\\
\eql -\hat{\cN}^{1/3}(2\Re\vev{\Sgm})\int\dr^4\tht\;\abs{\phi}^2
\brc{3\brkt{\cT+\bar{\cT}}-\brkt{a\cdot\cP}_{IJ}\tl{\cV}^I\tl{\cV}^J
+\cO(\hat{\Phi}^2,\tl{\cV}^3)}, \label{5DL_expand}
\eea
where $\tl{\cV}^I=-\der_y V^I+\tl{\Sgm}^I+\bar{\tl{\Sgm}}$
denotes the fluctuation part of $\cV^I$,
and $\cP^I_{\;\;J}$ is a projection operator defined as
\be
\cP^I_{\;\;J}(\cX) \equiv \dlt^I_{\;\:J}-\frac{\cX^I\cN_J}{3\cN}(\cX),
\label{def:cP}
\ee
which has a property,
\be
\brc{\cN_I\cP^I_{\;\;J}}(\cX) = \cP^I_{\;\;J}(\cX)\cX^J = 0.
\ee
The argument of $\brkt{a\cdot\cP}_{IJ}\equiv a_{IK}\cP^K_{\;\;J}$ is
$(2\Re\vev{\Sgm})$.
In the second line of (\ref{5DL_expand}), we can see that
the first term has the no-scale structure peculiar
to the 5D radion multiplet~\cite{Marti:2001iw},
and the second term does not include $\cT$ due to the projection
operator~$\cP^I_{\;\;J}$.
Hence $\Sgm^I$ are divided into two categories according to their origins.
\ignore{
In order to reproduce this result from (\ref{5D_action}),
we need to consider the replacement~(\ref{FDpromote}).
Then the gauge field~$A_\mu$ for $U(1)_A$,\footnote{
$U(1)_A$ is an automorphism of the $N=1$ superconformal
algebra~\cite{4Doffshell}. }
which is a vector component of $E_W$, is restored in the action
by promoting $\der_\mu$ to the covariant derivative~$\cD_\mu$.
This gauge field does not have a kinetic term in the action,
and should be treated as an auxiliary field.
After integrating it out, the correct coefficients~(\ref{def:a_IJ})
are reproduced.
}
\subsection{Gauging and mass scales}
In SUGRA, an introduction of any mass scales into the action requires
gauging some of the isometries on the hyperscalar manifold
by the moduli multiplets~$\bdm{V}^{I'}$, \ie,
we have to deal with the gauged SUGRA.
For example, the bulk cosmological constant is induced
when the compensator hypermultiplet~$(\Phi^1,\Phi^2)$ is charged,
and a bulk mass parameter for a physical hypermultiplet
is generated when it is charged for $\bdm{V}^{I'}$.
The gauging by $\bdm{V}^{I''}$ leads to the usual gauging
by a 4D massless gauge multiplet in 4D effective theory.
We omit the latter type of gauging in the following expressions
because it does not play a significant role in the derivation
of the effective theory and can be easily
restored in the 4D effective action.
We assume that all the gaugings by $\bdm{V}^{I'}$
are abelian, and are chosen to $\sgm_3$-direction
in the $(\Phi^{2a-1},\Phi^{2a})$-space, for simplicity.
Thus, the generators and the gauge couplings are chosen as
\be
\brkt{ig_{I'}t_{I'}}^{\hat{a}}_{\;\;\hat{b}}
= \sgm_3\otimes\diag\brkt{\frac{3}{2}k_{I'},c_{2I'},c_{3I'},
\cdots,c_{(n_H+1)I'}},
\ee
where $\sgm_3$ acts on each hypermultiplet~$(\Phi^{2a-1},\Phi^{2a})$.
Note that these coupling constants are $Z_2$-odd.
Such kink-type couplings can be realized in SUGRA context
by the mechanism proposed in Ref.~\cite{Bergshoeff:2000zn}.
In contrast to Ref.~\cite{Abe:2008an}, the compensator multiplet
also has non-vanishing charges,
which lead to the warping of the 5D spacetime.
Then, after rescaling chiral multiplets by a factor~$e^{3\sgm/2}$,
we obtain
\bea
\cL \eql -3\int\dr^4\tht\;\cN^{1/3}(\cV)\left\{
e^{-3k\cdot V}\abs{\Phi^1}^2+e^{3k\cdot V}\abs{\Phi^2}^2 \right. \nonumber\\
&&\hspace{30mm} \left.
-\sum_{a=2}^{n_H+1}\brkt{e^{-2c_a\cdot V}\abs{\Phi^{2a-1}}^2
+e^{2c_a\cdot V}\abs{\Phi^{2a}}^2}\right\}^{2/3} \nonumber\\
&&-2\sbk{\int\dr^2\tht\;\brc{
\Phi^1\brkt{\der_y+3k\cdot\Sgm}\Phi^2
-\sum_{a=2}^{n_H+1}\Phi^{2a-1}\brkt{\der_y+2c_a\cdot\Sgm}\Phi^{2a}}+\hc}
\nonumber\\
&&+\cL_{\rm vec}
+2\sum_{y_*=0,L}\sbk{\int\dr^2\tht\;\brkt{\Phi^2}^2
W^{(y_*)}\brkt{\hat{\Phi}^{2a}}+\hc}\dlt(y-y_*), \label{5D_action2}
\eea
where $k\cdot V\equiv\sum_{I'}k_{I'}V^{I'}$,
$c_a\cdot V\equiv\sum_{I'}c_{aI'}V^{I'}$,
and $\hat{\Phi}^{2a}=\Phi^{2a}/\Phi^2$.
For simplicity, we have introduced only superpotentials~$W^{(y_*)}$
in the boundary Lagrangians.
The hypermultiplets appear in $W^{(y_*)}$ only through $\hat{\Phi}^{2a}$
because physical chiral multiplets must have zero Weyl weights
in $N=1$ superconformal formulation~\cite{4Doffshell} and
$\Phi^{2a-1}$ vanish at the boundaries due to their orbifold parities.
The boundary Lagrangians can also depend on
boundary-localized 4D superfields, which are not considered
in this article.
The power of $\Phi^2$ in front of $W^{(y_*)}$ is determined by
the requirement that the argument of the $F$-term action formula
must have the Weyl weight 3.
(The Weyl weight of $\Phi^{\hat{a}}$ is $3/2$.)
\section{4D effective theory and its properties} \label{4Deffective}
\subsection{4D effective Lagrangian}
Following the off-shell dimensional reduction developed in Ref.~\cite{Abe:2007},
we can derive the 4D effective action, keeping the $N=1$ off-shell structure.
A detailed derivation is summarized in Appendix~\ref{derive:L_eff}.
The result is
\bea
\cL_{\rm eff} \eql -\frac{1}{4}\sbk{\int\dr^2\tht\;
\sum_r f_{\rm eff}^r(T)\tr\brkt{\cW^r\cW^r}+\hc} \nonumber\\
&&+\int\dr^4\tht\;\abs{\phi}^2\Omg_{\rm eff}\brkt{\abs{Q}^2,\Re T}
+\sbk{\int\dr^2\tht\;\phi^3 W_{\rm eff}(Q,T)+\hc},
\label{L_eff}
\eea
where the gauge multiplets are summarized in the matrix forms
for the non-abelian gauge groups,
the index~$r$ indicates the gauge sectors, and
$\cW^r$ is the field strength supermultiplet for
a massless 4D vector multiplet~$V^r$.
$Q_a$ ($a\geq 2$) and $T^{I'}$ are the zero-modes
for $\hat{\Phi}^{2a}$ and $\Sgm^{I'}$, respectively.
We have used the same symbols for the zero-modes~$V^r$ as
the corresponding 5D multiplets.
Each function in (\ref{L_eff}) is expressed as
\bea
f_{\rm eff}^r(T) \eql \sum_{I'}\xi^r_{I'}T^{I'}, \nonumber\\
\Omg_{\rm eff}\brkt{\abs{Q}^2,\Re T}
\defa -3e^{-K_{\rm eff}/3} \nonumber\\
\eql \hat{\cN}^{1/3}(\Re T)\left[
-3Y(k\cdot T)+2\sum_a Y((k+d_a)\cdot T)\abs{Q_a}^2 \right. \nonumber\\
&&\hspace{12mm}\left.
+\sum_{a,b}\tl{\Omg}^{(4)}_{a,b}(\Re T)\abs{Q_a}^2\abs{Q_b}^2
+\cO\brkt{(k\cdot\cP)^2}+\cO\brkt{\abs{Q}^6} \right], \nonumber\\
W_{\rm eff}(Q,T) \eql W^{(0)}(Q)+e^{-3k\cdot T}
W^{(L)}\brkt{e^{-d_a\cdot T}Q_a},
\label{defining_fcns}
\eea
where $\xi_{I'}^r$ are real constants determined from $C_{I'J''K''}$,
$d_{aI'}\equiv c_{aI'}-\frac{3}{2}k_{I'}$, and
\be
Y(z) \equiv \frac{1-e^{-2\Re z}}{2\Re z}. \label{def:Y}
\ee
The functions~$\tl{\Omg}^{(4)}_{a,b}$ are defined as
\be
\tl{\Omg}^{(4)}_{a,b} \equiv -\frac{\brkt{d_a\cdot\cP a^{-1}\cdot d_b}
\brc{Y((k+d_a+d_b)\cdot T)-\frac{Y(d_a\cdot T)Y(d_b\cdot T)}{Y(-k\cdot T)}}}
{\brc{(k+d_a)\cdot\Re T}\brc{(k+d_b)\cdot\Re T}}
+\frac{Y((k+d_a+d_b)\cdot T)}{3}. \label{tlOmg4}
\ee
In the derivation of $\Omg_{\rm eff}$ summarized
in Appendix~(\ref{derive:K_eff}), we have assumed that
\be
k_{I'}\cP^{I'}_{\;\;J'}(\Re T) = 0, \label{moduli_align}
\ee
in order to obtain an analytic expression.
Thus we focus on a case that the moduli VEVs are (at least approximately)
aligned to satisfy (\ref{moduli_align}) by some mechanism.
When the number of the moduli is two, the above effective Lagrangian
reduces to that in Ref.~\cite{Abe:2008an}
in the limit of $k_{I'}\to 0$.\footnote{
There are typos in Ref.~\cite{Abe:2008an}.
The indices of the derivatives of the norm functions should be replaced as
$\hat{\cN}_1\exch\hat{\cN}_2$ and $\hat{\cN}_{11}\exch\hat{\cN}_{22}$
in (2.15) and Sec.3.1 of Ref.~\cite{Abe:2008an}. }
The constraint~(\ref{moduli_align}) disappears in this limit.
\subsection{Superconformal gauge-fixing and mass dimension}
Here we mention the superconformal gauge-fixing in 4D SUGRA.
Since the effective action~(\ref{L_eff})
(with the promotion~(\ref{FDpromote}))
has the $N=1$ superconformal symmetries,
we have to impose the gauge-fixing conditions
in order to obtain the usual Poincar\'{e} SUGRA.
The extra symmetries to eliminate are
the dilatation~$\bdm{D}$, the $U(1)_A$ automorphism,
the conformal SUSY~$\bdm{S}$, and the conformal boost~$\bdm{K}$.
For a real general multiplet~$U=C+\tht\zeta+\bar{\tht}\bar{\zeta}+\cdots
+\frac{1}{2}\tht^2\bar{\tht}^2D$,
the $D$-term action formula is written as~\cite{Kugo:2002js,4Doffshell}
\be
\sbk{U}_D = e^{(4)}\brc{D+\frac{1}{3}C\brkt{
\cR^{(4)}-4\ep^{\mu\nu\rho\tau}\bar{\psi}_\mu\bar{\sgm}_\nu
\der_\rho\psi_\tau}
+\frac{4i}{3}\zeta\sgm^{\mu\nu}\der_\mu\psi_\nu+\cdots},
\ee
where $e^{(4)}$ is the determinant of the vierbein,
$\cR^{(4)}$ is the 4D Ricci scalar,
and $\sgm^{\mu\nu}\equiv\frac{1}{4}\brkt{\sgm^\mu\bar{\sgm}^\nu
-\sgm^\nu\bar{\sgm}^\mu}$.\footnote{
We follow the spinor notation of Ref.~\cite{Wess:1992cp}. }
We omitted the spacetime integral.
A conventional choice of the gauge-fixing is as follows.
The $\bdm{D}$-gauge is fixed so that the Einstein-Hilbert term is realized,
and the $\bdm{S}$-gauge is fixed so that the kinetic mixing between
the matter fermions and the gravitino is absent.
The corresponding gauge-fixing conditions are given by
\be
\left.\brkt{\abs{\phi}^2\Omg_{\rm eff}}\right|_0 = -3M_{\rm Pl}^2, \;\;\;\;\;
\left.\brkt{\abs{\phi}^2\Omg_{\rm eff}}\right|_\tht = 0. \label{GF_cond:4D}
\ee
The $U(1)_A$-symmetry is eliminated by fixing the phase of $\phi|_0$ to zero.
The $\bdm{K}$-gauge fixing condition is irrelevant to the current discussion.
\ignore{
Namely,
\bea
\phi|_0 \eql M_{\rm Pl}\brc{\hat{\cN}^{1/3}(\Re T)Y(k\cdot T)}^{-1/2}
+\cO(\abs{Q_a}^2),
\nonumber\\
\phi|_\tht \eql
\eea
}
The resultant Lagrangian in the gravitational sector is obtained as
\be
\cL_{\rm gravi} = -\frac{M_{\rm Pl}^2}{2}e^{(4)}\brkt{\cR^{(4)}
-4\ep^{\mu\nu\rho\tau}\bar{\psi}_\mu\bar{\sgm}_\tau\der_\nu\psi_\rho}
+\cdots. \label{L_gravi}
\ee
We comment on the relations to the 5D gauge-fixing we chose
in (\ref{GF_cond}).
From the definitions of $T^{I'}$ and $\phi$ in Appendix~\ref{derive:fW},
the $\bdm{D}$-gauge fixing in (\ref{GF_cond}) corresponds to
the following conditions in 4D effective theory.
\bea
\hat{\cN}^{1/3}(\Re\vev{T}) \eql \int_0^L\dr y\;\cN^{1/3}(2\Re\vev{\Sgm})
= M_5L, \nonumber\\
\vev{\phi} \eql \brkt{\vev{\Phi^2}}^{2/3} \simeq M_5. \label{rel:GF_cond}
\eea
We have assumed that VEVs of all the physical hypermultiplets
are much smaller than $M_{\rm Pl}^{3/2}$,
and used the facts that $\vev{\Sgm^{I''}}=0$ and
all $\vev{\Sgm^{I'}}$ have the same $y$-dependence
under the condition~(\ref{moduli_align}).
Namely we can read off the 5D scales~$M_5$ and $L^{-1}$
from (\ref{rel:GF_cond}).
Before the gauge fixing, all quantities in the action do not have
the mass dimension.
It can be defined after the $\bdm{D}$-gauge fixing
that introduces the mass scale into the theory.
Since $\bdm{D}$ corresponds to the scale transformation,
the mass dimension seems to be identified with
the $\bdm{D}$ charge, \ie, the Weyl weight.
However, they are completely different as shown in Table~\ref{mass_dimension}.
\begin{table}[t]
\begin{center}
\begin{tabular}{c||c|c|c|c|c|c|c}
\rule[-2mm]{0mm}{7mm} &
$x^\mu$ & $\tht^\alp$ & $e_\mu^{\;\;\underline{\nu}}$
& $\psi_\mu^\alp$ & $Q_a$ & $T^{I'}$ & $V^{I''}$ \\ \hline
\rule[-2mm]{0mm}{7mm} Weyl weight &
0 & $(-1/2)$ & $-1$ & $-1/2$ & 0 & 0 & 0 \\ \hline
\rule[-2mm]{0mm}{7mm} mass dimension &
$-1$ & $-1/2$ & 0 & 1/2 & 1 & $-1$ & 0
\\ \hline
\end{tabular}
\end{center}
\caption{The Weyl weights and the mass dimensions of
the coordinates and fields in 4D effective theory. }
\label{mass_dimension}
\end{table}
For instance, the former assigns a nonzero value to the coordinate~$x^\mu$
while the latter does not.
The mass dimensions of the gravitational fields are determined from
(\ref{L_gravi}).
As for the chiral and vector multiplets, the numbers in the table
denote those for the lowest components.
They increase for higher components by $1/2$.
The mass dimension of the moduli~$T^{I'}$ is determined
so that their VEVs have a dimension of length.
Note that the gauge-fixing conditions break the $N=1$ off-shell structure
and the theory cannot be expressed in terms of superfields any longer.
In order to express the action in terms of the component fields
that have the mass dimensions listed in Table~\ref{mass_dimension},
we rescale each quantity as
\bea
x^\mu & \to & M_{\rm Pl}x^\mu, \nonumber\\
\brkt{e_\mu^{\;\;\underline{\nu}}, \; \psi_\mu^\alp} & \to &
\frac{1}{M_{\rm Pl}}\brkt{e_\mu^{\;\;\underline{\nu}}, \;
\psi_\mu^\alp}, \nonumber\\
\brkt{Q_a, \; T^{I'}, \; V^{I''}} & \to &
\brkt{\frac{Q_a}{M_{\rm Pl}}, \; M_{\rm Pl}T^{I'}, \; V^{I''}}. \nonumber\\
\eea
Then the coupling constants~$k_{I'}$ and $c_{I'}$ are accompanied
with $M_{\rm Pl}$ in the rescaled action while $g_{I''}$ are not.
So we also rescale these constants as
\be
M_{\rm Pl}k_{I'}, \;\; M_{\rm Pl}c_{I'}, \;\; g_{I''} \;\;
\to \;\; k_{I'}, \;\; c_{I'}, \;\; g_{I''}.
\ee
Hence the moduli couplings~$k_{I'}$ and $c_{I'}$ are regarded as
mass parameters while the gauge couplings~$g_{I''}$ are dimensionless constants.
After this procedure,
$M_{\rm Pl}$ appears only in (\ref{L_gravi}) and the gravitational interactions.
\ignore{
\begin{enumerate}
\item Rescale the coordinates as
\be
\brkt{x^\mu, \;\; \tht^\alp} \to
\brkt{M_{\rm Pl}x^\mu, \;\; M_{\rm Pl}^{1/2}\tht^\alp}.
\label{rescale:cd}
\ee
\item Rescale the superfields as
\be
\brkt{\phi, \;\; Q_a, \;\; T^{I'}, \;\; V^{I''}} \to
\brkt{M_{\rm Pl}\phi, \;\; \frac{Q_a}{M_{\rm Pl}},
\;\; M_{\rm Pl}T^{I'}, \;\; V^{I''}},
\label{rescale:fd}
\ee
so that the mass dimensions of $\phi$, $Q_a$, $T^{I'}$ and $V^{I''}$
become 0, 1, $-1$, 0, respectively.
\item Rescale the coupling constants as
\be
\brkt{k_{I'}, \;\; c_{I'}, \;\; g_{I''}}
\to \brkt{\frac{k_{I'}}{M_{\rm Pl}}, \;\; \frac{c_{I'}}{M_{\rm Pl}},
\;\; g_{I''}}. \label{rescale:cp}
\ee
Then, the coupling constants for $\bdm{V}^{I'}$, $k_{I'}$ and $c_{I'}$,
are regarded as the mass parameters
while those for $\bdm{V}^{I''}$, $g_{I''}$,
are the usual dimensionless gauge coupling constants.
\end{enumerate}
Each component in a superfield is also rescaled in order to
absorb $M_{\rm Pl}$ induced by (\ref{rescale:cd}).
After this procedure,
$M_{\rm Pl}$ appears only in the gravitational interactions.
}
\subsection{Moduli kinetic terms}
Since the moduli VEVs satisfy (\ref{moduli_align}),
the combination~$k\cdot T$ is expressed as
\be
k\cdot T = \kp T_{\rm rad},
\ee
where
\be
\kp \equiv \frac{1}{L}\brkt{k\cdot\Re\vev{T}}, \;\;\;\;\;
T_{\rm rad} \equiv
L\frac{\hat{\cN}_{I'}}{3\hat{\cN}}(\Re\vev{T})T^{I'}.
\label{def:Trad}
\ee
We have determined the coefficient so that $\Re\vev{T_{\rm rad}}=L$.
Note that $T_{\rm rad}$ is identified with the zero-mode
of the 5D radion multiplet~$\cT$ defined in (\ref{def:cT}).
Since $\hat{\cN}^{1/3}(\Re T)$ is expanded around $T=\vev{T}$ as
\bea
\hat{\cN}^{1/3}(\Re(\vev{T}+\dlt T))
\eql \hat{\cN}^{1/3}(\Re\vev{T})\brc{
1+\frac{\hat{\cN}_{I'}}{3\hat{\cN}}(\Re\vev{T})
\Re\dlt T^{I'}+\cO(\dlt T^2)}
\nonumber\\
\eql \frac{1}{L}\hat{\cN}^{1/3}(\Re\vev{T})\Re T_{\rm rad}+\cO(\dlt T^2),
\eea
$\Omg_{\rm eff}$ in (\ref{defining_fcns}) becomes
\bea
\Omg_{\rm eff} \eql -\frac{3}{L}\hat{\cN}^{1/3}(\Re\vev{T})
\frac{1-e^{-2\kp\Re T_{\rm rad}}}{2\kp}+\cO(\dlt T^2)+\cdots,
\eea
where the ellipsis denotes terms involving other multiplets
than $T_{\rm rad}$.
This is the kinetic term of the radion multiplet
in the Randall-Sundrum spacetime~\cite{Luty:2000ec}.
Thus, the alignment of moduli VEVs in (\ref{moduli_align})
is interpreted as the condition
for the spacetime to be the Randall-Sundrum spacetime,
and $\kp$ defined in (\ref{def:Trad}) is identified with
the AdS curvature scale that is related to the bulk cosmological constant.
After the gauge-fixing~(\ref{GF_cond:4D}), we obtain
\be
M_{\rm Pl}^2 = -\frac{1}{3}\vev{\abs{\phi}^2\Omg_{\rm eff}}
\simeq -\frac{1}{3}M_5^2\cdot
\brkt{-\frac{3}{L}}\cdot M_5L \cdot \frac{1-e^{-2\kp L}}{2\kp}
= \frac{M_5^3\brkt{1-e^{-2\kp L}}}{2\kp}. \label{rel:Mpl}
\ee
We have used (\ref{rel:GF_cond}).
Eq.(\ref{rel:Mpl}) is the well-known relation
in the Randall-Sundrum spacetime~\cite{RS}.
The other moduli are orthogonal to $T_{\rm rad}$ in the moduli space,
When there exist $n_V$ moduli, they are parametrized
by the coordinate system~$\brc{\vph^i}$
($i=1,\cdots,n_V-1$) on the $(n_V-1)$-dimensional submanifold determined
by~\footnote{
One choice of $\brc{\vph^i}$ is $\vph^i\equiv T^{I'=i}-\vev{T^{I'=i}}$.
In this case, $T^{n_V}$ becomes a function of $\vph^i$ through (\ref{VSmanifold}).
}
\be
\hat{\cN}(\Re T(\vph))=\hat{\cN}(\Re\vev{T}).
\label{VSmanifold}
\ee
Since $\hat{\cN}_{I'}(\Re T)\frac{\der T^{I'}}{\der\vph^i}=0$,
$(k\cdot\cP)_{I'}$ is expressed in terms of $\vph^i$ as
\be
\brkt{k\cdot\cP}_{I'} = -\brc{\frac{(k\cdot\Re T)\hat{\cN}_{I'J'}}
{3\hat{\cN}}}_{\vph=0}
\Re\brkt{\left.\frac{\der T^{J'}}{\der\vph^i}\right|_{\vph=0}\vph^i}
+\cO(\vph^2).
\ee
The kinetic terms for $\vph^i$ are contained in the $\cO((k\cdot\cP)^2)$-terms
in (\ref{defining_fcns}), which start from~\footnote{
This can be calculated by the perturbative expansion of
the first equation in (\ref{dg/dU})
in terms of $\abs{k}\equiv(\sum_{I'}k_{I'}^2)^{1/2}$.
}
\be
\dlt\Omg_{\rm eff} = -\frac{1}{4}\brc{\hat{\cN}^{1/3}
\brkt{k\cdot\cP}_{I'} a^{I'J'}\brkt{k\cdot\cP}_{J'}}(\Re T)
+\cO\brkt{k^4,\abs{Q}^2}.
\ee
\subsection{Quadratic terms in $\bdm{\Omg_{\rm eff}}$ and Yukawa hierarchy}
\label{Yukawa_hierarchy}
The coefficients of $\abs{Q_a}^2$ in $\Omg_{\rm eff}$ are important
for generating the fermion mass hierarchy.
The Yukawa couplings can be introduced only in the boundary actions
due to the $N=2$ SUSY in the bulk.
We assume that they are contained in $W^{(0)}$ at $y=0$,
\be
W^{(0)} = \sum_{a,b,c}\lmd_{abc}\hat{\Phi}^{2a}\hat{\Phi}^{2b}\hat{\Phi}^{2c}
+\cdots,
\ee
where $\lmd_{abc}$ are the holomorphic Yukawa coupling constants
and are supposed to be of $\cO(1)$.
Then the effective theory has the Yukawa couplings,
\be
W_{\rm eff} = \sum_{a,b,c}\lmd_{abc}Q_a Q_b Q_c+\cdots.
\ee
The physical Yukawa couplings~$y_{abc}$ are obtained
by the canonical normalization of the chiral superfields~$Q_a$, and we have
\be
y_{abc} = \frac{\lmd_{abc}}{\sqrt{\vev{Y_aY_bY_c}}},
\label{yabc}
\ee
where
\be
Y_a \equiv 2\hat{\cN}^{1/3}(\Re T)
\brc{Y((k+d_a)\cdot T)+\tl{\Omg}_{a,X}^{(4)}(\Re T)\abs{X}^2+\cO(\abs{X}^4)}.
\label{def:Y_a}
\ee
The function~$Y(z)$ is always positive,
and approximated as
\be
Y(z) \simeq \begin{cases} \frac{1}{2\Re z}, & \Re z>0 \\
\frac{1}{2\abs{\Re z}}\exp\brc{2\abs{\Re z}}. & \Re z<0 \end{cases}
\ee
From the 5D viewpoint, the wave function of $Q_a$ is localized
toward $y=0$ ($y=L$)
in the case that $(k+d_a)\cdot\vev{\Re T}$ is positive (negative).
As we can see from (\ref{yabc}), $y_{abc}$ is of $\cO(1)$
when all the relevant fields are localized toward $y=0$,
while it is exponentially small when there is a field localized toward $y=L$
among them.
This is the well-known split fermion mechanism~\cite{Arkani-Schmaltz}.
\subsection{Quartic couplings in $\bdm{\Omg_{\rm eff}}$
and soft SUSY-breaking masses} \label{quartic_cp}
The coefficients of $\abs{Q_a}^2\abs{Q_b}^2$ have a peculiar form
to 5D SUGRA.
This type of terms are important because they lead to
the soft SUSY-breaking masses
for the sfermions when $Q_a$ and $Q_b$ are identified with
the quark or lepton superfield and the SUSY-breaking superfield~$X$,
respectively.
Notice that the first term in (\ref{tlOmg4}) is absent
in the single modulus case due to the projection operator~$\cP^{I'}_{\;\;J'}$.
It is induced by integrating out the vector multiplets~$V^{I'}$,
which are the $N=2$ partners of the moduli
multiplets~$\Sgm^{I'}$~\cite{Abe:2008an}.
\begin{figure}[t,b]
\centering \leavevmode
\includegraphics[width=70mm]{Feyn1.eps}
\caption{Feynmann diagrams contributing to $\tl{\Omg}_{a,X}^{(4)}$
in the multi moduli case.
}
\label{diagram}
\end{figure}
The relevant Feynmann diagrams are depicted in Fig.~\ref{diagram}.
This can be seen from the fact that the coefficient
of the first term in (\ref{tlOmg4}) involves the inverse matrix of $a_{I'J'}$,
which comes from the propagator for $V^{I'}$.
The existence of the projection operator~$\cP^{I'}_{\;\;J'}$ indicates that
the graviphoton multiplet~$V_B$ defined in (\ref{def:V_B}) does not
contribute to $\tl{\Omg}_{a,b}^{(4)}$.
This can be understood from the fact that
most of the components of $V_B$ are auxiliary fields,
as mentioned in Sec.~\ref{sc_GF}.
In the next section, we will consider a case that
the $F$-term of one chiral multiplet~$X$ in the effective theory
provides the main source of SUSY breaking and $\vev{X}\simeq 0$.
In such a case, the $\abs{Q_a}^2\abs{X}^2$-term contributes
to the soft SUSY-breaking mass for the scalar component of $Q_a$,
and it is expressed as
(see (\ref{sp}) in Sec.~\ref{visible})
\be
m_a^2 \simeq -\abs{F^X}^2\frac{\tl{\Omg}^{(4)}_{a,X}(\Re\vev{T})}
{Y((k+d_a)\cdot\vev{T})}. \label{ap_m_soft}
\ee
Let us first consider the single modulus case.
In this case, $\tl{\Omg}^{(4)}_{a,X}$ is always positive
because the first term in (\ref{tlOmg4}) is absent.
Thus the soft scalar masses in (\ref{ap_m_soft}) become
\be
m_a^2 \simeq -\abs{F^X}^2\frac{Y((\kp+d_a+d_X)L)}
{3Y((\kp+d_a)L)}, \label{m_soft:single}
\ee
and are found to be tachyonic.
These tachyonic masses can be saved by quantum effects
in some cases.
The soft masses in (\ref{m_soft:single}) are exponentially suppressed
when $\kp+d_a<0$ and $\kp+d_X>0$.
This corresponds to a case that the matter~$Q_a$ is localized
around $y=L$ while $X$ is around $y=0$.
In such a case, quantum effects to the soft scalar masses become dominant
and may lead to non-tachyonic masses.
However the large top quark mass cannot be realized
because the top Yukawa coupling is suppressed in that case.
(Recall that the Yukawa couplings are localized at the $y=0$ boundary.)
This problem can be evaded in the multi moduli case.
Let us consider a case that
\be
d_a\cdot\Re\vev{T} < -\kp L < 0 < d_X\cdot\Re\vev{T}.
\label{rel:dTs}
\ee
In this case, the $y=0$ boundary is identified with the UV brane, and
$Q_a$ and $X$ are localized around the IR and UV branes respectively
since $(k+d_a)\cdot\Re\vev{T}<0$ and $(k+d_X)\cdot\Re\vev{T}>0$.
VEV of $\tl{\Omg}_{a,b}^{(4)}$ is approximately expressed as
\bea
\tl{\Omg}_{a,X}^{(4)}(\Re\vev{T}) \sma
\frac{d_a\cdot\cP a^{-1}\cdot d_X}
{\brc{(k+d_a)\cdot\Re\vev{T}}\brc{(k+d_X)\cdot\Re\vev{T}}}
\frac{Y(d_a\cdot\vev{T})Y(d_X\cdot\vev{T})}{Y(-\kp L)} \nonumber\\
\sma \frac{d_a\cdot\cP a^{-1}\cdot d_X}
{\brc{(k+d_a)\cdot\Re\vev{T}}\brc{(k+d_X)\cdot\Re\vev{T}}}\cdot
\frac{-\kp L e^{-2(k+d_a)\cdot\Re\vev{T}}}
{2\brkt{d_a\cdot\Re\vev{T}}\brkt{d_X\cdot\Re\vev{T}}}. \nonumber\\
\label{ap_tlOmg4}
\eea
Therefore, (\ref{ap_m_soft}) becomes
\be
m_a^2 \simeq -\abs{F^X}^2\frac{(d_a\cdot\cP a^{-1}\cdot d_X)\kp L}
{\brkt{d_a\cdot\Re\vev{T}}\brkt{d_X\cdot\Re\vev{T}}
\brc{(k+d_X)\cdot\Re\vev{T}}}. \label{ap:soft_mass}
\ee
The sign of $m_a^2$ now depends on
the (truncated) norm function~$\hat{\cN}$ and the directions of the gauging
for $V^{I'}$.
In fact, we can always realize non-tachyonic soft masses
for any choices of $\hat{\cN}$
by choosing the directions of the gauging such that
\be
\frac{d_a\cdot\cP a^{-1}\cdot d_X}
{\brkt{d_a\cdot\Re\vev{T}}\brkt{d_X\cdot\Re\vev{T}}} < 0.
\label{cond:non-tachyonic}
\ee
Furthermore, if $n_V$-dimensional vectors~$\vec{d}_{a}$
point to the same direction,
\be
\vec{d}_a\propto\vec{n}, \label{cond:d_a}
\ee
the soft masses~$m_a^2$ become independent of the ``flavor index''~$a$.
(The direction~$\vec{n}$ must not be parallel to $\vec{k}$,
otherwise all $m_a^2$ vanish.)
This opens up the possibility to solve the SUSY flavor problem.
We will discuss this issue in the next section.
\ignore{
Finally we provide an interpretation of the above flavor universality
from the 5D viewpoint.
Recall that the dominant part in $\tl{\Omg}_{aX}^{(4)}$
under the condition~(\ref{rel:dts}) originates from the diagram
in Fig.~\ref{diagram}.
The internal line represents the $n$-th KK modes of $V^{I'}$,
which are integrated out.
The effective couplings~$g_a^{(n)}$ and $g_X^{(n)}$ are defined as
}
A similar result is also obtained in a case that
\be
d_a\cdot\Re\vev{T} < 0 < -\kp L < d_X\cdot\Re\vev{T}.
\ee
Conditions for obtaining non-tachyonic (and flavor-universal) soft masses
are the same as (\ref{cond:non-tachyonic}) (and (\ref{cond:d_a})).
In this case, however, the $y=0$ boundary becomes the IR brane,
and the approximate expressions of the soft masses are suppressed
from (\ref{ap:soft_mass}) by a factor~$e^{-2\kp L}\gg 1$.\footnote{
The typical SUSY-breaking mass scale~$M_{\rm SB}$ is also suppressed
by the same factor in this case.
So a ratio~$m_a/M_{\rm SB}$ is not suppressed.
(See Sec.~\ref{visible}.)}
\section{Illustrative model} \label{specific_model}
Now we specify the model that realizes the set-up mentioned in Sec.~\ref{setup},
and show some phenomenological analysis.
\subsection{Hidden (mediation) sector contents} \label{hidden}
In this paper we assume that a single chiral multiplet~$X$ originating
from a 5D hypermultiplet is responsible for the spontaneous SUSY breaking.
We do not specify the potential of the hidden sector $X$ and moduli $T^{I'}$
chiral multiplets. There are various ways to stabilize the moduli,
including the size of the extra dimension.
In general, the mechanism for the moduli stabilization determines
the $F$-terms of $T^{I'}$, and thus affects the mediation of SUSY breaking
to the MSSM sector.
Here we do not specify the moduli stabilization and SUSY breaking mechanism,
and treat (VEVs of) $F^X$ and $F^{T^{I'}}$ as free parameters,\footnote{
We do not consider the $D$-term SUSY breaking in this article.}
while VEV of (the lowest component of) $X$ is assumed to be almost
vanishing $\langle X \rangle \ll \langle T^{I'} \rangle \simeq {\cal O}(1)$
in the 4D Planck mass unit.\footnote{
Concrete moduli stabilization and SUSY breaking mechanisms were studied
in our previous work~\cite{Abe:2008an} based on Ref.~\cite{Abe:2006xp},
which are also applicable here.}
Instead of the non-vanishing $F$-terms, $F^{T^{I'}}$ ($I'=1,2,3$) and $F^X$,
we use ratios of them~$\alp_{I'}$ and a typical scale of SUSY breaking~$M_{\rm SB}$
in order to parametrize the soft SUSY breaking parameters.
They are defined as
\bea
\alp_{I'} \defa \frac{\cF^{T^{I'}}}{\cF^X}, \;\;\;\;\;
M_{\rm SB} \equiv \left|\cF^X\right|, \nonumber\\
\cF^A \defa E^A_{\;\;B}F^B, \label{def:M_SB}
\eea
where $E^A_{\;\;B}$ is the vielbein\footnote{
We use the same index for the flat and curved coordinates
on the K\"ahler manifold to save characters. }
for the K\"{a}hler metric~$K_{A\bar{B}}\equiv\der_A\der_{\bar{B}}K_{\rm eff}$.
\ignore{
\be
\alp_{I'} \equiv \frac{\sqrt{K_{I'\bar{I'}}}
\left|F^{T^{I'}}\right|}
{\sqrt{K_{X\bar{X}}}\left|F^X\right|}, \;\;\;\;\;
M_{\rm SB} \equiv \sqrt{K_{X\bar{X}}}\left|F^X\right|,
\label{def:M_SB}
\ee
where $K_{I'\bar{I}'}$ and $K_{X\bar{X}}$ are the diagonal components
of the K\"ahler metric shown in (\ref{K-metric}).
}
Then the vacuum value of the scalar potential is written as
\bea
V \eql \sum_A\left|\cF^A\right|^2-3e^{K} \left| W \right|^2 \nonumber\\
\eql \left( 1+\sum_{I'=1}^3\abs{\alpha_{I'}}^2 \right) M_{\rm SB}^2 -3m_{3/2}^2,
\eea
where $m_{3/2}= \langle e^{K/2}|W| \rangle$ is the gravitino mass.
The moduli stabilization at a SUSY breaking Minkowski minimum
$\langle V \rangle \simeq 0$ required by the observation
leads to a relation
\begin{eqnarray}
m_{3/2} &\simeq& \sqrt{ \frac{1}{3}
\left( 1+\sum_{I'=1}^3\abs{\alpha_{I'}}^2 \right)}\,M_{\rm SB}.
\end{eqnarray}
The ratios~$\alp_{I'}$ parameterize contributions
of ``the moduli mediation'' induced by $F^{T^{I'}}$
compared to that of ``the direct mediation'' induced by $F^X$.
In the following analysis we mainly consider cases that
\begin{eqnarray}
\abs{\alpha_{I'}} \ = \ {\cal O}(1)
\ &\textrm{or}& \ {\cal O} \left( 1/(4 \pi^2) \right).
\end{eqnarray}
The latter values can be realized by the numerical value of
$1/\ln (M_{\rm Pl}/{\rm TeV})$ if the moduli $T^{I'}$
are stabilized by nonperturbative effects like gaugino
condensations at a SUSY anti-de Sitter vacuum, which is
uplifted to the almost Minkowski minimum by the vacuum energy of
the (TeV scale) SUSY breaking sector~\cite{Choi:2004sx,Choi:2005ge}.
We can identify $F^X$ with a source of the uplifting vacuum energy
based on a scenario of F-term uplifting~\cite{Abe:2006xp,Dudas:2006gr}.
For the phenomenological analysis,
we consider a case with three $Z_2$-odd $U(1)_{I'}$ vector multiplets
$\bdm{V}^{I'}$ in 5D, where $I'=1,2,3$.
They generate three moduli chiral multiplets
\be
T^{I'}=(T^1, T^2, T^3),
\ee
in the 4D effective theory
and we choose the (truncated) norm function as
\be
\hat{\cN}(\Re T) = (\Re T^1)(\Re T^2)(\Re T^3).
\label{nft1t2t3}
\ee
Then the matrix~$a_{I'J'}$ defined in (\ref{def:a_IJ}) becomes diagonal.
Explicit forms of the K\"ahler metric~$K_{A\bar{B}}$ and $a_{I'J'}$ are
shown in Appendix~\ref{explicit_forms}.
In the flat case ($\kp L=0$), the K\"ahler metric also becomes diagonal.
For concreteness, we further assume the moduli VEVs as
\be
\brkt{\Re\vev{T^1},\Re\vev{T^2},\Re\vev{T^3}} = (4,4,1). \;\;\;\;\;
\mbox{(in the $M_{\rm Pl}$ unit)} \label{VEV:assumption}
\ee
Then the condition to realize the Randall-Sundrum spacetime~(\ref{moduli_align})
fixes the gauging direction of the compensator multiplets as
\be
(k_1,k_2,k_3) = \brkt{\frac{1}{4},\frac{1}{4},1}k_3.
\label{cpst_gauging}
\ee
The value of $k_3$ is determined by the warp factor
through $\kp L=k\cdot\Re\vev{T}=3k_3$.
\subsection{Visible sector contents and soft SUSY-breaking parameters}
\label{visible}
We assume that the visible sector consists of
the following MSSM matter contents:
\bea
(V_1,V_2,V_3) &:& \mbox{gauge vector multiplets}, \nonumber\\
(\cQ_i,\cU_i,\cD_i) &:& \mbox{quark chiral multiplets}, \nonumber\\
(\cL_i,\cE_i) &:& \mbox{lepton chiral multiplets}, \nonumber\\
(\cH_u,\cH_d) &:& \mbox{Higgs chiral multiplets},
\eea
where $V_1$, $V_2$, $V_3$ denote
the gauge multiplets for $U(1)_Y,SU(2)_L,SU(3)_C$
originating from $\bdm{V}^{I''}$,
and the others are chiral multiplets from 5D hypermultiplets.
The index~$i=1,2,3$ denotes the generation.
We also assume an approximate global~$U(1)_R$-symmetry that is responsible
for the dynamical SUSY breaking~\cite{Nelson:1993nf}.
We assign the R-charge as shown in Table~\ref{Rcharge}.
\begin{table}[t]
\begin{center}
\begin{tabular}{c||c|c|c} \hline
\rule[-2mm]{0mm}{7mm}
Multiplet & $\cQ_i$, $\cU_i$, $\cD_i$, $\cL_i$, $\cE_i$ &
$\cH_u$, $\cH_d$ & $X$ \\ \hline
\rule[-2mm]{0mm}{7mm} R-charge & 1/2 & 1 & 2 \\ \hline
\end{tabular}
\end{center}
\caption{$U(1)_R$-charges of the chiral multiplets. }
\label{Rcharge}
\end{table}
It is supposed to be broken by the nonperturbative effects.
Then, the holomorphic Yukawa couplings and the $\mu$-term
in the boundary superpotentials as well as the gauge kinetic functions
are independent of $X$.
We further assume that these terms exist only
at the $y=0$ boundary.
The resulting gauge kinetic functions and the superpotential
for the visible sector in the 4D effective theory are parametrized as
\bea
f^r_{\rm eff}(T) \eql \sum_{I'}\xi^r_{I'}T^{I'}, \nonumber\\
W_{\rm MSSM} \eql \zeta_0\cH_u\cH_d+\lmd_{ij}^u\cH_u\cQ_i\cU_j
+\lmd_{ij}^d\cH_d\cQ_i\cD_j+\lmd_{ij}^e\cH_d\cL_i\cE_j,
\label{wmssm}
\eea
where $r=1,2,3$ for $U(1)_Y$, $SU(2)_L$, $SU(3)_C$,
and $\xi^r_{I'}$ are real constants determined by
the coefficients $C_{I'J''K''}$ in the norm function ${\cal N}$,
while $\zeta_0$ and $\lmd_{ij}^{u,d,e}$ are in general complex constants.
After the canonical normalization, the $\mu$ parameter is expressed as
\be
\mu = \Lvev{\frac{\zeta_0}{\sqrt{\hat{\cN}^{1/3}(\Re T) Y(k\cdot T)
Y_{\cH_u}Y_{\cH_d}}}}
= \Lvev{\frac{\zeta_0}{\sqrt{2Y(\kp L)Y_{\cH_u}Y_{\cH_d}}}},
\label{expr:mu}
\ee
where $Y_a$ is defined in (\ref{def:Y_a}),
and we have used (\ref{VEV:assumption}) at the second equality.
The physical Yukawa couplings are expressed as
\be
y_{ij}^u = \frac{\lmd_{ij}^u}
{\sqrt{\vev{Y_{\cH_u}Y_{\cQ_i}Y_{\cU_j}}}}, \;\;\;\;\;
y_{ij}^d = \frac{\lmd_{ij}^d}
{\sqrt{\vev{Y_{\cH_d}Y_{\cQ_i}Y_{\cD_j}}}}, \;\;\;\;\;
y_{ij}^e = \frac{\lmd_{ij}^e}
{\sqrt{\vev{Y_{\cH_d}Y_{\cL_i}Y_{\cE_j}}}}.
\ee
The holomorphic Yukawa couplings~$\lmd_{ij}^x$ ($x=u,d,e$)
are assumed to be of $\cO(1)$.
The hierarchical structure of the Yukawa couplings are obtained
by choosing the moduli couplings~$c_{a I'}$
as explained in Sec.~\ref{Yukawa_hierarchy}.
The soft SUSY-breaking parameters in the MSSM are defined by
\bea
\cL_{\rm soft} \eql -\sum_{Q_a}m_{Q_a}^2\abs{Q_a}^2
-\frac{1}{2}\sum_r M_r\tr\brkt{\lmd^r\lmd^r} \\
&&-\sum_{i,j}\brc{B\mu H_u H_d+y_{i,j}^u A_{i,j}^u H_u \tl{q}_i \tl{u}_j
+y_{i,j}^d A_{i,j}^d H_d\tl{q}_i\tl{d}_j
+y_{i,j}^e A_{i,j}^e H_d\tl{l}_i\tl{e}_j+\hc}, \nonumber
\eea
where $\lmd^r$ ($r=1,2,3$) and $Q_a=H_{u,d}$, $\tl{q}_i$, $\tl{u}_i$, $\tl{d}_i$,
$\tl{l}_i$, $\tl{e}_i$ are the gauginos and
the scalar components of $\cH_{u,d}$, $\cQ_i$, $\cU_i$, $\cD_i$,
$\cL_i$, $\cE_i$, respectively.
The fields are all {\it canonically normalized}.
These parameters are induced
by the formulae~\cite{Choi:2005ge,Kaplunovsky:1993rd},
\bea
M_r \eql \vev{F^A\der_A\ln\brkt{\Re f^r_{\rm eff}}}, \;\;\;\;\;
m_{Q_a}^2 =
-\vev{F^A\bar{F}^{\bar{B}}\der_A\der_{\bar{B}}\ln Y_{Q_a}}, \nonumber\\
B \eql -\Lvev{F^A\der_A\ln\brkt{\frac{\zeta_0}{\hat{\cN}^{1/3}(\Re T)
Y(k\cdot T)Y_{\cH_u}Y_{\cH_d}}}}-m_{3/2}e^{i\vph}, \nonumber\\
A_{ij}^u \eql \vev{F^A\der_A\ln\brkt{Y_{\cH_u}Y_{\cQ_i}Y_{\cU_j}}}, \;\;\;\;\;
A_{ij}^d = \vev{F^A\der_A\ln\brkt{Y_{\cH_d}Y_{\cQ_i}Y_{\cD_j}}}, \nonumber\\
A_{ij}^e \eql \vev{F^A\der_A\ln\brkt{Y_{\cH_d}Y_{\cL_i}Y_{\cE_j}}},
\label{sp}
\eea
where $A,B=X,T^1,T^2,T^3$, and $\vph\equiv\arg\vev{W}$.
Now we show the (squared) soft scalar masses~$m_{Q_a}^2$ as functions
of the charges for the $Z_2$-odd vector multiplets~$V^{I'}$ ($I'=1,2,3$).
Since $\vev{Y_a}$ in the above formulae are functions of
$(k+d_a)\cdot\Re\vev{T}$,
we normalize each charge as
\begin{eqnarray}
\tilde{c}^{I'}_a \defa (k_{I'}+d_{aI'})\Re\vev{T^{I'}}
= (c_{aI'}-k_{I'}/2) \Re\vev{T^{I'}}, \label{def:tlc}
\end{eqnarray}
without summations for the index~$I'$.
The soft scalar mass~$m_{Q_a}^2$ varies exponentially over $\cO(1)$ ranges of $\tl{c}_a^{I'}$.
Thus we plot a quantity defined as
\begin{eqnarray}
\bt(m_{Q_a}^2) &\equiv&
\log_{10} \frac{\sqrt{|m_{Q_a}^2|}}{M_{\rm SB}}.
\label{hatmqa}
\end{eqnarray}
In Figs.~\ref{cdsmsq1} and \ref{cdsmsq2}, we assume that
$\alp_1=\alp_3=1/(4\pi^2)$, $\alp_2=2/(4\pi^2)$.
Namely, contributions of the moduli mediation are tiny.
In Fig.~\ref{cdsmsq3}, we assume that $\alp_1=\alp_2=1$, $\alp_3=2$,
\ie, contributions of the moduli mediation are comparable to
that of the direct mediation.
The charge assignment of $X$ is chosen as
$(\tl{c}_X^1,\tl{c}_X^2,\tl{c}_X^3)=(\tl{c}_X-\frac{2\kp L}{3},
\frac{\kp L}{3},\frac{\kp L}{3})$ in all figures,
while that of $Q_a$ is chosen as
$(\tl{c}^1_{Q_a},\tl{c}^2_{Q_a},\tl{c}^3_{Q_a})
=(\frac{\kp L}{3},\tl{c}_a-\frac{2\kp L}{3},\frac{\kp L}{3})$
in Figs.~\ref{cdsmsq1} and \ref{cdsmsq3}, and
$(\tl{c}^1_{Q_a},\tl{c}^2_{Q_a},\tl{c}^3_{Q_a})
=(\tl{c}_a-\frac{2\kp L}{3},\frac{\kp L}{3},\frac{\kp L}{3})$ in Fig.~\ref{cdsmsq2}.
The surface with (without) a mesh describes a region~$m_{Q_a}^2>0$
($m_{Q_a}^2<0$).
\begin{figure}[t]
\centering \leavevmode
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{softmass1N.eps}
(a) $\kp L = -3.6$
\end{center}
\end{minipage}
\hfill
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{softmass1zero.eps}
(b) $\kp L = 0$
\end{center}
\end{minipage}
\hfill
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{softmass1P.eps}
(c) $\kp L = 3.6$
\end{center}
\end{minipage}
\caption{
The charge dependences of $\bt(m_{Q_a}^2)$ defined in
Eq.(\ref{hatmqa}) with the norm function (\ref{nft1t2t3})
and $\alpha_1=\alp_3=1/(4\pi^2)$, $\alp_2=2/(4\pi^2)$.
The charge assignment for $Q_a$ and $X$ is chosen as
$(\tl{c}_{Q_a}^1,\tl{c}_{Q_a}^2,\tl{c}_{Q_a}^3)
=(\frac{\kp L}{3},\tl{c}_a-\frac{2\kp L}{3},\frac{\kp L}{3})$
and $(\tl{c}_X^1,\tl{c}_X^2,\tl{c}_X^3)
=(\tl{c}_X-\frac{2\kp L}{3},\frac{\kp L}{3},\frac{\kp L}{3})$.
The surface with (without) a mesh
describes the region $m_{Q_a}^2>0$ ($m_{Q_a}^2<0$).
}
\label{cdsmsq1}
\end{figure}
\begin{figure}[t]
\centering \leavevmode
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{softmass2N.eps}
(a) $\kp L = -3.6$
\end{center}
\end{minipage}
\hfill
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{softmass2zero.eps}
(b) $\kp L = 0$
\end{center}
\end{minipage}
\hfill
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{softmass2P.eps}
(c) $\kp L = 3.6$
\end{center}
\end{minipage}
\caption{
The charge dependences of $\bt(m_{Q_a}^2)$ defined in
Eq.(\ref{hatmqa}) with the norm function (\ref{nft1t2t3})
and $\alpha_1=\alp_3=1/(4\pi^2)$, $\alp_2=2/(4\pi^2)$.
The charge assignment for $Q_a$ and $X$ is chosen as
$(\tl{c}_{Q_a}^1,\tl{c}_{Q_a}^2,\tl{c}_{Q_a}^3)
=(\tl{c}_a-\frac{2\kp L}{3},\frac{\kp L}{3},\frac{\kp L}{3})$
and $(\tl{c}_X^1,\tl{c}_X^2,\tl{c}_X^3)
=(\tl{c}_X-\frac{2\kp L}{3},\frac{\kp L}{3},\frac{\kp L}{3})$.
The surface with (without) a mesh
describes the region $m_{Q_a}^2>0$ ($m_{Q_a}^2<0$).
}
\label{cdsmsq2}
\end{figure}
\begin{figure}[t]
\centering \leavevmode
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{smass3N.eps}
(a) $\kp L = -3.6$
\end{center}
\end{minipage}
\hfill
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{smass3zero.eps}
(b) $\kp L = 0$
\end{center}
\end{minipage}
\hfill
\begin{minipage}{0.3\linewidth}
\begin{center}
\includegraphics[width=\linewidth]{smass3P.eps}
(c) $\kp L = 3.6$
\end{center}
\end{minipage}
\caption{
The charge dependences of $\bt(m_{Q_a}^2)$ defined in
Eq.(\ref{hatmqa}) with the norm function (\ref{nft1t2t3})
and $\alp_1=\alp_2=1$, $\alp_3=2$.
The charge assignment for $Q_a$ and $X$ is chosen as
$(\tl{c}_{Q_a}^1,\tl{c}_{Q_a}^2,\tl{c}_{Q_a}^3)
=(\frac{\kp L}{3},\tl{c}_a-\frac{2\kp L}{3},\frac{\kp L}{3})$
and $(\tl{c}_X^1,\tl{c}_X^2,\tl{c}_X^3)
=(\tl{c}_X-\frac{2\kp L}{3},\frac{\kp L}{3},\frac{\kp L}{3})$.
The surface with (without) a mesh
describes the region $m_{Q_a}^2>0$ ($m_{Q_a}^2<0$).
}
\label{cdsmsq3}
\end{figure}
Note that the above charge assignments satisfy (\ref{cond:d_a}) because
they are rewritten as
\bea
\brkt{d_X^1,d_X^2,d_X^3} \eql \brkt{d_X,0,0}, \nonumber\\
\brkt{d_{Q_a}^1,d_{Q_a}^2,d_{Q_a}^3} \eql \begin{cases} \displaystyle
\brkt{0,d_a,0}, & (\mbox{in Figs.~\ref{cdsmsq1} and
\ref{cdsmsq3}}) \\ \displaystyle
\brkt{d_a,0,0}. & (\mbox{in Fig.~\ref{cdsmsq2}}) \end{cases}
\eea
where $d_X\equiv(\tl{c}_X-\kp L)/4$ and $d_a\equiv(\tl{c}_a-\kp L)/4$.
Let us first consider a case that $\alp_{I'}\ll 1$, \ie,
the direct mediation dominates.
The soft scalar mass is expressed from (\ref{ap_m_soft}) as
\be
\frac{m_{Q_a}^2}{M_{\rm SB}^2} \simeq
-\frac{Y(\kp L)\tl{\Omg}^{(4)}_{Q_a,X}(\Re\vev{T})}{2Y(\tl{c}_a)Y(\tl{c}_X)}.
\label{ratio:m/M}
\ee
We can see that $\abs{m_{Q_a}^2}>M_{\rm SB}^2$ for $\tl{c}_a\tl{c}_X\gg 1$
in Figs.~\ref{cdsmsq1} and \ref{cdsmsq2}.
This behavior can be understood from the fact that
$Q_a$ and $X$ localize toward the same boundary in such a region.
Especially in the warped case, the soft scalar mass is enhanced
when they localize toward the IR boundary.
On the other hand, they localize toward the opposite boundaries
in a region~$\tl{c}_a\tl{c}_X<-1$.
Notice that $m_{Q_a}^2/M_{\rm SB}^2$ is not exponentially suppressed
and remains to be of $\cO(1)$ in this region,
because they are not sequestered
due to the existence of the contact terms~$\abs{Q_a}^2\abs{X}^2$
in $\Omg_{\rm eff}$ induced by integrating out the heavy $Z_2$-odd vector
multiplets~$V^{I'}$.
This is in sharp contrast to the single modulus case
where the sequestering occurs.
Especially, in a region that
$\tl{c}_a\tl{c}_X<-1$ and $(\tl{c}_a-\kp L)(\tl{c}_X-\kp L)< -1$,
such induced contact terms dominate,
and (\ref{ratio:m/M}) can be approximated as
\be
\frac{m_{Q_a}^2}{M_{\rm SB}^2} \simeq
-\frac{e^{2\kp L}Y(\kp L)}{2Y(-\kp L)}
\frac{d_{Q_a}\cdot\cP a^{-1}\cdot d_X}
{(d_{Q_a}\cdot\Re\vev{T})(d_X\cdot\Re\vev{T})}
\simeq -\frac{d_{Q_a}\cdot\cP a^{-1}\cdot d_X}
{2(d_{Q_a}\cdot\Re\vev{T})(d_X\cdot\Re\vev{T})}.
\label{ap:m/M}
\ee
Since the charge assignments of $Q_a$ satisfy
the condition~(\ref{cond:d_a}), the flavor dependence of (\ref{ap:m/M}) cancels.
With our choice of the norm function and the assumption~(\ref{VEV:assumption}),
its approximate value is $16/3$ ($-32/3$)
in the case of Fig.~\ref{cdsmsq1} (Fig.~\ref{cdsmsq2}),
irrespective of the value of $\kp L$.
The non-tachyonic condition in such a flavor-universal
region~(\ref{cond:non-tachyonic}) is satisfied
for the charge assignment of Fig.~\ref{cdsmsq1},
while not for that of Fig.~\ref{cdsmsq2},
as we can see the figures.
In a case that contributions from the moduli mediation are not negligible,
\ie, $\abs{\alp_{I'}}=\cO(1)$, the above behaviors are disturbed.
The expression of the soft scalar mass in such a case
is given in (\ref{explicit:ma/MSB}).
In the flat case ($\kp L=0$), for example, it is written as
\be
\frac{m_{Q_a}^2}{M_{\rm SB}^2} \simeq
-\frac{\tl{\Omg}^{(4)}_{Q_a,X}(\Re\vev{T})}{2Y(\tl{c}_a)Y(\tl{c}_X)}
+\sum_{I'}\frac{\abs{\alp_{I'}}^2}{3}
-\sum_{I',J'}\alp_{I'}\bar{\alp}_{J'}
\tl{c}_{Q_a}^{I'}\tl{c}_{Q_a}^{J'}\cY(\tl{c}_a),
\ee
where $\cY(x)$ is a function defined in (\ref{def:cY}).
We have used the specific form of the norm function~(\ref{nft1t2t3}).
In spite of the nontrivial $\tl{c}_a$-dependence of the third term,
there is still a region in which $m_{Q_a}^2$ is almost flavor-universal.
This is due to the property of the function~$\cY(x)$
that $x^2\cY(x)\simeq 1$ for $\abs{x}\simgt 3$.
\ignore{
we find
$|m_{Q_a}^2| \gtrsim M_{\rm SB}^2$
($|m_{Q_a}^2| \sim M_{\rm SB}^2$)
for
$\tilde{c}^a \tilde{c}^X \gtrsim 1$
($\tilde{c}^a \tilde{c}^X \lesssim 1$)
because in this case $Q_a$ and $X$ localize toward
the same (opposite) fixed point to each other.
This behavior can be understood from the geometrical point of view.
For the warped case $\kp L > 0$ ($\kp L < 0$)
in Figs.~\ref{cdsmsq1} and \ref{cdsmsq2},
the absolute value of scalar mass squared
$|m_{Q_a}^2|$ is enhanced (suppressed) for the region
$\tilde{c}_a$, $\tilde{c}_X < 0$.
This is the effect of the warp factor $e^{\kp L}$.
The sign of the scalar mass squared $m_{Q_a}^2$
(with the same warp factor) is flipped between
Fig.~\ref{cdsmsq1} and Fig.~\ref{cdsmsq2}
in the region $|\tilde{c}_a \tilde{c}_X| \gtrsim 1$.
Therefore, the tachyonic soft scalar mass can be avoided for
a suitable charge assignment even with certain fixed values of
$\tilde{c}_a$ and $\tilde{c}_X$. This is one of the most important
property with multiple moduli in order to realize realistic Yukawa
hierarchies ({\it i.e.} observed quark/lepton masses and mixings)
without tachyonic squarks/sleptons.
}
By making use of the above properties,
we construct a realistic model in the next subsection,
and analyze the flavor structure of fermions and
sfermions as well as the other phenomenological features.
We comment that the boundary induced K\"{a}hler potentials~$K^{(y_*)}$
($y_*=0,L$) are neglected in this paper.
They may disturb the flavor structure
if they dominate the contributions from the bulk.
Here we just assume that such boundary contributions are small enough
compared to those from the bulk.
\subsection{Phenomenological analysis}
In the following phenomenological analysis,
the warp factor is chosen as
\be
\kp L = 3.6, \label{kL}
\ee
so that $e^{\kp L}=\cO(M_{\rm Pl}/M_{\rm GUT})$.
This determines the compensator charges in (\ref{cpst_gauging}) as $k_3=1.2$.
The typical KK mass scale is set to the GUT scale,
\be
M_{\rm KK} \equiv \frac{\kp\pi}{e^{\kp L}-1}
= M_{\rm GUT}. \label{mKK}
\ee
This determines $\kp$ (and $L$) from (\ref{kL}).
The 4D effective theory is valid below $M_{\rm KK}$.
Here we comment on a consistency condition of our 5D setup.
In order for the 5D description of the theory to be valid,
the 5D curvature~$\cR^{(5)}$ must satisfy
the condition~$\abs{\cR^{(5)}}<M_5^2$~\cite{Davoudiasl:1999tf}.
For the Randall-Sundrum spacetime, $\cR^{(5)}=-20\kappa^2$.
Thus, together with the relation~(\ref{rel:Mpl})
and the definition of $M_{\rm KK}$, the consistency condition is rewritten as
\be
e^{\kp L} < \frac{\sqrt{2}\pi}{20^{3/4}}\frac{M_{\rm Pl}}{M_{\rm KK}}
\simeq 0.47\frac{M_{\rm Pl}}{M_{\rm KK}}.
\ee
For our choice of $M_{\rm KK}$ in (\ref{mKK}), this indicates
that $\kp L<4.0$, which is satisfied by (\ref{kL}).
In order to realize phenomenologically viable fermion and
sfermion flavor structures, we assign the following
$U(1)_{I'}$ charges for the $Z_2$-odd vector multiplets~$V^{I'}$
($I'=1,2,3$) to the MSSM matter multiplets
$Q_a=({\cQ_i,\cU_i,\cD_i,\cL_i,\cE_i,\cH_u,\cH_d})$.
For the values of
${\rm Re}\,\langle T^{I'} \rangle$ and $k_{I'}$
given by (\ref{VEV:assumption}) and (\ref{cpst_gauging}),
the $U(1)_1$ charges are chosen as
\begin{eqnarray}
&\tl{c}^{I'=1}_{\cQ_i} = (1.2, 1.2, 0.5), \quad
\tl{c}^{I'=1}_{\cU_i} = (1.2, 1.2, 0.5), \quad
\tl{c}^{I'=1}_{\cD_i} = (1.2, 1.2, 1.2), & \nonumber \\
&\tl{c}^{I'=1}_{\cL_i} = (1.2, 1.2, 1.2), \quad
\tl{c}^{I'=1}_{\cE_i} = (1.2, 1.2, 1.2), & \nonumber \\
&\tl{c}^{I'=1}_{\cH_u} = 1.0, \quad
\tl{c}^{I'=1}_{\cH_d} = 1.2, \quad
\tl{c}^{I'=1}_{X} = 8.7, &
\label{fc1}
\end{eqnarray}
the $U(1)_2$ charges are assigned as
\begin{eqnarray}
&\tl{c}^{I'=2}_{\cQ_i} = (-7.9, -5.9,0), \quad
\tl{c}^{I'=2}_{\cU_i} = (-10.4, -5.9,0), \quad
\tl{c}^{I'=2}_{\cD_i} = (-6.4, -6.9, -4.9),& \nonumber \\
&\tl{c}^{I'=2}_{\cL_i} = (-6.9, -6.9, -4.9), \quad
\tl{c}^{I'=2}_{\cE_i} = (-9.4, -3.9, -3.9),& \nonumber \\
&\tl{c}^{I'=2}_{\cH_u} = 0, \quad
\tl{c}^{I'=2}_{\cH_d} = -3.4, \quad
\tl{c}^{I'=2}_{X} = 1.2, &
\label{fc2}
\end{eqnarray}
and the $U(1)_3$ charges are assigned as
\begin{eqnarray}
&\tl{c}^{I'=3}_{\cQ_i} = (1.2, 1.2, 0), \quad
\tl{c}^{I'=3}_{\cU_i} = (1.2, 1.2, 0), \quad
\tl{c}^{I'=3}_{\cD_i} = (1.2, 1.2, 1.2),& \nonumber \\
&\tl{c}^{I'=3}_{\cL_i} = (1.2, 1.2, 1.2), \quad
\tl{c}^{I'=3}_{\cE_i} = (1.2, 1.2, 1.2),& \nonumber \\
&\tl{c}^{I'=3}_{\cH_u} = 0, \quad
\tl{c}^{I'=3}_{\cH_d} = 1.2, \quad
\tl{c}^{I'=3}_{X} = 1.2, &
\label{fc3}
\end{eqnarray}
\ignore{
These charges are almost determined
uniquely by the requirement that the observed
quark and charged lepton masses and the absolute values
of CKM mixings are realized\footnote{Note that the completely
precise matching of the predicted and observed values shown in
Table~\ref{pqlmm} is meaningless because the former values
are calculated from the tree-level effective theory at
$M_{\rm GUT}$ with certain reference values of ${\cal O}(1)$
holomorphic Yukawa couplings $\lambda^{u,d,e}_{ij}$ and
some one-loop renormalization effects below $M_{\rm GUT}$.}
as shown in Table~\ref{pqlmm} with ${\cal O}(1)$ values of
the holomorphic Yukawa couplings $\lambda^{u,d,e}_{ij}$
in the superpotential (\ref{wmssm}).
}
These charges satisfy (\ref{cond:d_a})
for the first two generations of quark and lepton multiplets.
With this charge assignment,
the observed quark and charged lepton masses and the absolute values
of CKM mixings are realized, as shown in Table~\ref{pqlmm},
with ${\cal O}(1)$ values of
the holomorphic Yukawa couplings $\lambda^{u,d,e}_{ij}$
in the superpotential (\ref{wmssm}).
\begin{table}[t]
\begin{center}
\begin{tabular}{|c||c|c|} \hline
& Predicted & Observed \\ \hline
$(m_u, m_c, m_t)/m_t$ &
$(1.4 \times 10^{-5}, 7.38 \times 10^{-3}, 1.0)$ &
$(1.5 \times 10^{-5}, 7.37 \times 10^{-3}, 1.0)$ \\ \hline
$(m_d, m_s, m_b)/m_b$ &
$(1.2 \times 10^{-3}, 2.41 \times 10^{-2}, 1.0)$ &
$(1.2 \times 10^{-3}, 2.54 \times 10^{-2}, 1.0)$ \\ \hline
$(m_e, m_\mu, m_\tau)/m_\tau$ &
$(2.871 \times 10^{-4}, 5.955 \times 10^{-2}, 1.0)$ &
$(2.871 \times 10^{-4}, 5.959 \times 10^{-2}, 1.0)$ \\ \hline
$|V_{CKM}|$ &
\begin{minipage}{0.38\linewidth}
\begin{eqnarray}
\left(
\begin{array}{ccc}
0.97324 & 0.2298 & 0.00337 \\
0.2297 & 0.97235 & 0.042 \\
0.00637 & 0.0417 & 0.999112
\end{array}
\right)
\nonumber
\end{eqnarray} \\*[-20pt]
\end{minipage}
&
\begin{minipage}{0.38\linewidth}
\begin{eqnarray}
\left(
\begin{array}{ccc}
0.97428 & 0.2253 & 0.00347 \\
0.2252 & 0.97345 & 0.041 \\
0.00862 & 0.0403 & 0.999152
\end{array}
\right)
\nonumber
\end{eqnarray} \\*[-20pt]
\end{minipage}
\\ \hline
\end{tabular}
\end{center}
\caption{The predicted quark and charged lepton masses
as well as the absolute values of CKM mixings compared
with the experimental data~\cite{Nakamura:2010zzi}.
The flavor charges are chosen as shown in
Eqs.(\ref{fc1}) and (\ref{fc2}).}
\label{pqlmm}
\end{table}
After fixing all the $U(1)_{I'}$ charges, the remaining parameters
are the coefficients $\xi^r_{I'}$ in the effective
gauge kinetic functions $f^r_{\rm eff}(T)$ in (\ref{wmssm}).
One of $\xi^r_{I'}$ for each $r$ is determined by matching $f^r_{\rm eff}(\vev{T})$
with the observed values of the SM gauge couplings, \ie,
by the condition for the gauge coupling unification at $M_{\rm GUT}$.
The remaining $\xi^r_{I'}$ control the gaugino masses~$M_r$ at $M_{\rm GUT}$.
In the following analysis, all the gauge couplings, Yukawa couplings,
the $\mu$-term and soft SUSY breaking parameters in the visible sector
are evaluated at the EW scale~$M_{\rm EW}$ by using the 1-loop renormalization
group (RG) equations of MSSM, where we neglect effects of all
the Yukawa couplings except for the top Yukawa coupling.
In order to estimate the SUSY flavor violations,
we rotate the soft scalar mass matrices
$(m^2_{\tl{f}_{L,R}})_{ij}=\diag\brkt{m^2_{\tl{f}_{L,R}1},m^2_{\tl{f}_{L,R}2},
m^2_{\tl{f}_{L,R}3}}$
and the scalar trilinear coupling matrices
$(\tilde{A}^f)_{ij}=(y^f)_{ij}(A^f)_{ij}$
into the super-CKM basis and describe them as
\be
\hat{m}^2_{\tilde{f}_L} = (V^f_L)^\dagger m^2_{\tilde{f}_L} V^f_L, \quad
\hat{m}^2_{\tilde{f}_R} = (V^f_R)^\dagger m^2_{\tilde{f}_R} V^f_R, \quad
\hat{A}^f = (V^f_L)^\dagger \tilde{A}^f V^f_R,
\ee
where $f=u,d,e$ and $V^u_L=V^d_L\equiv V^q_L$.
The unitary matrices $V^f_{L,R}$ are defined by
\be
(V^f_L)^\dagger y_f V^f_R = \frac{1}{v_f}\diag\brkt{m_{f1},m_{f2},m_{f3}},
\ee
where $v_f=(\sin \beta, \cos \beta, \cos \beta)v$ and $v \simeq 174$ GeV.
Here we consider a case of $\tan\bt=4$ as an example.\footnote{
In fact, $\tan\bt$ is not a free parameter in our original setup.
One way to treat $\tan\bt$ as a free parameter
is to introduce another $\mu$-term, $\zeta_L\cH_u\cH_d$, on the $y=L$ boundary.
Then the constant~$\zeta_0$ in (\ref{expr:mu}) and (\ref{sp}) is replaced by
$\zeta\equiv\zeta_0+e^{-(c_{\cH_u}+c_{\cH_d})\cdot T}\zeta_L$, and
we can control the value of $\tan\bt$ by varying $\zeta_L$
keeping $\vev{\zeta}$ unchanged.
}
Then we define mass insertion parameters as~\cite{Misiak:1997ei}
\begin{eqnarray}
&\displaystyle
(\delta_{LL}^f)_{ij} =
\frac{(\hat{m}^2_{\hat{f}_L})_{ij}
+\left( \left( m_f \right)_i - \rho_{LL}^f \right)
\delta_{ij}
}{
\sqrt{
(\hat{m}^2_{\tilde{f}_L})_{ii}
(\hat{m}^2_{\tilde{f}_L})_{jj}}}, \quad
(\delta_{RR}^f)_{ij} =
\frac{(\hat{m}^2_{\hat{f}_R})_{ij}
+\left( \left( m_f \right)_i - \rho_{RR}^f \right)
\delta_{ij}
}{
\sqrt{
(\hat{m}^2_{\tilde{f}_R})_{ii}
(\hat{m}^2_{\tilde{f}_R})_{jj}}},&
\nonumber \\
&\displaystyle
(\delta_{LR}^f)_{ij} =
\frac{v_f (\hat{A}^f)_{ij}
-\mu_f \left( m_f \right)_i \delta_{ij}
}{
\sqrt{
(\hat{m}^2_{\tilde{f}_L})_{ii}
(\hat{m}^2_{\tilde{f}_R})_{jj}}}
\ = \ (\delta_{RL}^f)_{ji}^\ast,&
\end{eqnarray}
where
\begin{eqnarray}
\mu_f &=& \left(
\cot \beta, \tan \beta, \tan \beta \right) \mu,
\nonumber \\
\rho_{LL}^f &=&
\frac{\cos 2 \beta}{6} \left(
M_Z^2-4M_W^2,\, M_Z^2+2M_W^2,\, -3M_Z^2+6M_W^2 \right),
\nonumber \\
\rho_{RR}^f &=&
\frac{\cos 2 \beta}{3} \sin^2 \theta_W \left(
-2M_Z^2,\, M_Z^2,\, 3M_Z^2 \right),
\nonumber \\
M_Z &\simeq& 91.2 \ {\rm GeV}, \quad
M_W \ \simeq \ 80.1 \ {\rm GeV}, \quad
\sin^2 \theta_W \ \simeq \ 0.23,
\end{eqnarray}
and $\mu$ is determined
by the minimization condition of the Higgs potential.
\begin{figure}[t]
\centering \leavevmode
\hfill
\includegraphics[width=0.35\linewidth]{1t2w.eps}
\hfill
\includegraphics[width=0.35\linewidth]{1t2s.eps}
\hfill
\includegraphics[width=0.25\linewidth]{1t2l.eps}
\hfill
\caption{
Contours of
$|(\delta_{LR}^e)_{21}|$,
$|(\delta_{LL}^d)_{23}(\delta_{LR}^d)_{33}|$,
$|\Delta_{\mu}|$ and
$m_{H_0}$
as functions of the gaugino mass ratios
$r_1=M_1/M_3$ and $r_2=M_2/M_3$ at $M_{\rm GUT}$.
The region $3 \le r_2 \le 6$
is magnified in the right panel.
The parameters are chosen as $M_{\rm SB}=100$~GeV,
$\alpha_1=1,\alpha_2=1/2$, $\alpha_3=1/(4\pi^2)$, and $\tan \beta =4$.
The gluino mass~$M_3$ is set to 343~GeV at $M_{\rm GUT}$.
}
\label{1t2}
\end{figure}
\begin{figure}[t]
\centering \leavevmode
\hfill
\includegraphics[width=0.35\linewidth]{pt2w.eps}
\hfill
\includegraphics[width=0.35\linewidth]{pt2s.eps}
\hfill
\includegraphics[width=0.25\linewidth]{pt2l.eps}
\hfill
\caption{
Contours of
$|(\delta_{LR}^e)_{21}|$,
$|(\delta_{LL}^d)_{23}(\delta_{LR}^d)_{33}|$,
$|\Delta_{\mu}|$ and
$m_{H_0}$
as functions of the gaugino mass ratios
$r_1=M_1/M_3$ and $r_2=M_2/M_3$ at $M_{\rm GUT}$.
The region $3 \le r_2 \le 6$
is magnified in the right panel.
The parameters are chosen as $M_{\rm SB}=200$~GeV,
$\alpha_1=1/2$, $\alpha_2=1/(8\pi^2)$, $\alpha_3=1/(4\pi^2)$,
and $\tan \beta =4$.
The gluino mass~$M_3$ is set to 383~GeV at $M_{\rm GUT}$.
}
\label{pt2}
\end{figure}
\begin{figure}[t]
\centering \leavevmode
\hfill
\includegraphics[width=0.35\linewidth]{ptiw.eps}
\hfill
\includegraphics[width=0.35\linewidth]{ptis.eps}
\hfill
\includegraphics[width=0.25\linewidth]{ptil.eps}
\hfill
\caption{
Contours of
$|(\delta_{LR}^e)_{21}|$,
$|(\delta_{LL}^d)_{23}(\delta_{LR}^d)_{33}|$,
$|\Delta_{\mu}|$ and
$m_{H_0}$
as functions of the gaugino mass ratios
$r_1=M_1/M_3$ and $r_2=M_2/M_3$ at $M_{\rm GUT}$.
The region $3 \le r_2 \le 6$
is magnified in the right panel.
The parameters are chosen as $M_{\rm SB}=500$~GeV,
$\alpha_1=1/(4\pi^2),\alpha_2=2/(4 \pi^2)$, $\alpha_3=1/(4 \pi^2)$,
and $\tan \beta =4$.
The gluino mass~$M_3$ is set to 416~GeV at $M_{\rm GUT}$.
}
\label{pti}
\end{figure}
Furthermore we introduce a quantity
\be
\Dlt_\mu \equiv \frac{\mu}{2M_Z^2}\frac{\der M_Z^2}{\der\mu},
\ee
which describes the sensitivity of the Z-boson mass $M_Z$
to the $\mu$ parameter~\cite{Barbieri:1987fn}.
In Figs.~\ref{1t2}, \ref{pt2} and \ref{pti},
we plot contours of various quantities as functions of
the gaugino mass ratios at $M_{\rm GUT}$,
\be
r_1 \equiv M_1/M_3, \;\;\;\;\;
r_2 \equiv M_2/M_3.
\ee
Contours of $|(\delta_{LR}^e)_{21}|$,
$|(\delta_{LL}^d)_{23}(\delta_{LR}^d)_{33}|$,
$|\Delta_{\mu}|$ and the lightest CP-even neutral Higgs mass~$m_{H_0}$
are shown in Fig.~\ref{1t2} for $M_{\rm SB}=100$~GeV,
$\alpha_1=1$, $\alpha_2=1/2$ and $\alpha_3=1/(4\pi^2)$.
Similar contours are plotted in Fig.~\ref{pt2}
for$M_{\rm SB}=200$~GeV, $\alpha_1=1/2$, $\alpha_2=1/(8\pi^2)$ and $\alpha_3=1/(4\pi^2)$, and
in Fig.~\ref{pti} for $M_{\rm SB}=500$ GeV,
$\alpha_1=\alpha_3=1/(4 \pi^2)$ and
$\alpha_2=2/(4 \pi^2)$.
The gluino mass~$M_3$ at $M_{\rm GUT}$ is 343~GeV, 383~GeV and
416~GeV in Figs.~\ref{1t2}, \ref{pt2} and \ref{pti}, respectively.
Ratios of the gaugino masses at $M_{\rm EW}$ to those at $M_{\rm GUT}$ are
given by
\be
\frac{M_1(M_{\rm EW})}{M_1(M_{\rm GUT})}=0.4, \;\;\;\;\;
\frac{M_2(M_{\rm EW})}{M_2(M_{\rm GUT})}=0.8, \;\;\;\;\;
\frac{M_3(M_{\rm EW})}{M_3(M_{\rm GUT})}=2.9.
\ee
The gaugino masses at $M_{\rm EW}$ are read off from these relations.
The curves with fixed values of $m_{H_0}$ represents
the upper bound on the lightest CP-even Higgs mass
in our model~\cite{Carena:1995wu}.
The region surrounded by the curve representing
$|\Delta_{\mu}| = 10$ is free from the little hierarchy problem of MSSM.
There is no (less than 10\%) fine-tuning of $\mu$
in this region in order to realize the correct EW symmetry breaking.
As for the SUSY flavor violations,
the most stringent experimental constraints on the
mass insertion parameters in our model are typically
expressed as
$|(\delta_{LR}^e)_{21}| \lesssim 10^{-7}$ and
$|(\delta_{LL}^d)_{23}(\delta_{LR}^d)_{33}| \lesssim 10^{-4}$
which come from the upper bound on the branching ratios of
$\mu \to e \gamma$ and $b \to s \gamma$, respectively~\cite{Abe:2004tq}.
Recall that contribution of the direct mediation by $F^X$
has only weak flavor dependence for our charge assignment
while that of the moduli mediation
by $F^{T^{I'}}$ can disturb such flavor structures.
Namely, the flavor dependence of the soft masses becomes weaker
when $\alp_{I'}\ll 1$.
We emphasize that the allowed region from $\mu \to e \gamma$
is much wider in Fig.~\ref{pt2} with $\alpha_{2}=1/(8 \pi^2)$,
compared with the one in Fig.~\ref{1t2} with $\alpha_{2}=1/2$.
This is because the flavor structure of the first and second
generations in the lepton sector are governed by the $U(1)_2$ charges,
and only $\alp_2$ affects the flavor structure.
Note that the contours of
$|(\delta_{LL}^d)_{23}(\delta_{LR}^d)_{33}|$ in Fig.~\ref{pti}
is drawn differently compared with Figs.~\ref{1t2} and \ref{pt2}.
This is because contributions of the direct mediation dominates
over those of the moduli mediation
since $\alp_{I'}\ll 1$ in Fig.~\ref{pti},
and then the scaling of the soft terms are changed.
It is commonly said that models with the gravity-mediated SUSY breaking
suffer from the SUSY flavor problem.
However,
we find that suitable charge assignments for $U(1)_{I'}$ do not
cause the SUSY flavor problem while realize a viable flavor structure for quarks
and leptons (without tachyonic squarks and sleptons).
We should emphasize that this is due to the existence of
multiple moduli, which induce additional contact terms~$\abs{Q_a}^2\abs{X}^2$
in 4D effective K\"ahler potential.
This is in sharp contrast to models with a single modulus
discussed in many papers.
Finally we comment that the Higgs mass bound seems to be most stringent
in Figs.~\ref{1t2}, \ref{pt2} and \ref{pti}. This is due to the fact that
the visible sector is assumed to be MSSM with $\tan \beta =4$.
It may become milder without affecting the flavor structure if we take
a larger value of $\tan \beta$~\footnote{
For large values of $\tan\beta$, contributions of the bottom Yukawa coupling
become important in the RG running, which we have neglected here.
}
and/or extend the Higgs sector, such as the next to minimal SUSY SM.
Since we are focusing on the flavor structure here,
we would leave analyses on such extensions for a separate paper.
Even for MSSM with $\tan \beta =4$,
we find a region where all the experimental constraints
considered here are satisfied in Fig.~\ref{pt2}.
\section{Summary} \label{summary}
We have systematically studied the SUSY flavor structure of
generic 5D SUGRA models, where all the hidden and visible
sector fields are living in the whole 5D bulk spacetime,
where $N=2$ SUSY exists.
In order to realize the observed quark and lepton masses
and mixings, the visible sector fields are quasi-localized
in the extra dimension by a suitable charge assignment
for $Z_2$-odd $U(1)_{I'}$ vector multiplets~$V^{I'}$.
This type of models have been considered in many papers,
but most of them assume that there is just a single
$Z_2$-odd vector multiplet, \ie, the graviphoton multiplet.
However, it has been shown in Ref.~\cite{Abe:2008ka}
that induced squark and slepton masses
become tachyonic in such a case.
Besides, too large flavor violation generically occurs
in the SUGRA models, \ie, the SUSY flavor problem.
In our previous work~\cite{Abe:2008an}, we pointed out a new possibility
of avoiding such problems by introducing
an extra $Z_2$-odd vector multiplet
other than the graviphoton multiplet.
In such a case, additional contributions to
4D effective K\"ahler potential~$K_{\rm eff}$
appear after integrating out the $Z_2$-odd vector multiplets,
and they affect the flavor structure of the soft SUSY-breaking mass matrices.
In this paper, we have extended our previous work to more generic cases
and specify conditions to solve the tachyonic sfermion
problem~(\ref{cond:non-tachyonic}) and
the SUSY flavor problem~(\ref{cond:d_a}).
In fact, through a detailed phenomenological analysis, we have
explicitly shown that the SUSY flavor problem can be avoided
by introducing multiple vector multiplets $V^{I'}$ without
encountering tachyonic sfermion problem mentioned above.
Therefore we conclude that the SUSY flavor structure of
gravity-mediated SUSY breaking scenario can be
{\it controllable}~\cite{Conlon:2007dw}
once it is concretely constructed, contrary to the general
criticism that SUSY flavor violation is problematic for it.
The additional contributions to $K_{\rm eff}$ by integrating out
the bulk SUGRA fields
have been discussed in the context of the string theory
in Ref.~\cite{Anisimov:2001zz}.
Because 5D SUGRA is the simplest set-up for the brane-world models,
we can derive an explicit form of $K_{\rm eff}$ directly from
the higher-dimensional theory.
This enables us to perform detailed analyses on it,
as we have done in this paper.
Our systematic analyses also owe to the existence of
the off-shell description of SUGRA~\cite{Kugo-Ohashi,Kugo:2002js}.
It makes the derivation of the 4D effective theory transparent
by utilizing the off-shell dimensional reduction~\cite{Abe:2007}.
The results obtained in this paper are quite generic
when the hierarchical flavor structure originates
from the wave function localization in the extra dimension.
Most of the results in Sec.~\ref{specific_model} do not much depend on
the choice of the norm function~(\ref{nft1t2t3}) if we choose
a suitable charge assignment for $V^{I'}$.
Finally we emphasize that the multi moduli case discussed in this paper
is naturally realized
when we consider the low-energy effective theories of the string theory.
\section*{Acknowledgments}
This work was supported in part by the Waseda University
Grant for Special Research Projects No.2011B-177 (H. A.),
and Grant-in-Aid for Young Scientists (B) No.22740187
from Japan Society for the Promotion of Science (Y. S.).
H. A. and Y. S. thank the Yukawa Institute for Theoretical
Physics at Kyoto University.
Discussions during the YITP workshop~"Summer Institute 2011"
were useful to complete this work.
|
\section{Introduction}
\label{sec:intro}
In the classification scheme proposed by \cite{1936rene.book.....H} galaxies are arranged on a sequence going from ellipticals to lenticulars and, from these, to spirals of progressively later type (Sa to Sc). Ellipticals are dominated by a stellar bulge while the spiral sequence is essentially one of decreasing bulge-to-disc ratio. The intermediate position of lenticular galaxies in this scheme has lead to the common idea that all early-type galaxies (ellipticals and lenticulars; hereafter ETGs) have higher bulge-to-disc ratio than spirals.
The other difference between ETGs and spirals is that the former lack the blue spiral arms typical of the latter \citep{1926ApJ....64..321H}. It was early recognised that this corresponds to a lack of star formation in ETGs, leading to the simplified picture that their stellar populations are uniformly old.
In contrast with this traditional view, ETGs exhibit a large variety of shapes and some authors suggest that their bulge-to-disc ratio can in fact be as low as that of Sc, disc-dominated spirals \citep{1951ApJ...113..413S,1970ApJ...160..831S,1976ApJ...206..883V}. Furthermore, following early insights by \cite{1981ApJ...249...48G} and \cite{1982ApJ...261...85R}, it is now established that a large fraction of ETGs are forming small amounts of stars or have done so in their recent past \citep[e.g.,][]{1993PhDT.......172G,2000AJ....119.1645T,2005ApJ...619L.111Y,2007ApJS..173..619K,2010MNRAS.404.1775T}.
Support to these ideas comes from recent studies of nearby ETGs using integral-field spectroscopy \citep{2002MNRAS.329..513D}. These show that most ETGs host a rotating, kinematically cold component \citep{2008MNRAS.390...93K} whose stars are usually younger and more metal-rich than those in the bulge \citep{2010MNRAS.408...97K}.
These results are placed on a firm statistical ground by ATLAS$^{\rm 3D}$, a multi-wavelength study of a volume-limited sample of 260 morphologically selected ETGs \citep[hereafter Paper I]{2011MNRAS.413..813C}. We find that as many as 80 percent of all ETGs in the nearby Universe consist of nearly axisymmetric, fast rotating stellar systems (\citealt[herafter Paper II]{2011MNRAS.414.2923K}; \citealt[hereafter Paper III]{2011MNRAS.414..888E}), most of which resemble spiral galaxies with the arms removed \citep[hereafter Paper VII]{2011MNRAS.tmp.1249C}.
The presence of a substantial disc and the occurrence of star formation in ETGs imply that cold gas has played an important role in their evolution. Indeed, \cite{2011MNRAS.tmp.1500K} suggest that most ETGs grow a stellar disc following gas cooling. In this respect, two fundamental lines of research are the direct observation of neutral hydrogen gas ({\sc H\,i}) and molecular gas (H$_2$, observed through the radiation emitted by CO molecules). In spirals, {\sc H\,i}\ is an essential constituent of the disc, and is the material from which H$_2$ and subsequently new stars form. Understanding the {\sc H\,i}\ and H$_2$ properties of ETGs is therefore crucial to investigate the origin of their structure and star formation history, how they continue to evolve at $z=0$, and why they are so different from spiral galaxies.
Results from the ATLAS$^{\rm 3D}$\ CO survey are presented in \cite[hereafter Paper IV]{2011MNRAS.414..940Y}. Here we present an {\sc H\,i}\ survey of 166 nearby ETGs belonging to the ATLAS$^{\rm 3D}$\ sample. Thanks to its unprecedented combination of sample size, sample selection, depth and resolution of the {\sc H\,i}\ data, and availability of multi-wavelength data this survey represents a significant improvement over previous studies.
The study of {\sc H\,i}\ in ETGs dates back to the end of the 1960's. Early work already showed that ETGs have lower $M$(\hi)$/L_\mathrm{B}$ than spirals, consistent with their redder colour \citep[e.g.,][]{1969A&A.....3..281G,1970A&A.....6..453B,1972A&A....21..303B,1972AJ.....77..568G,1973A&A....22..281B,1975ApJ...202....7G,1976ApJ...209..710B,1977A&A....60..361B,1977AJ.....82..106K,1978ApJ...222..800K,1979ApJ...234..448K,1979A&A....75....7B,1979ApJ...227..776K,1980ApJ...242..931S,1982A&AS...50..451R}. Based on these data, \cite{1985AJ.....90..454K} and \cite{1986AJ.....91...23W} analysed the {\sc H\,i}\ content of $\sim150$ ellipticals and $\sim300$ lenticulars, respectively. They detected {\sc H\,i}\ in about 15 percent of all ellipticals and 25 percent of all lenticulars for a typical $M$(\hi)\ detection limit of a few times $10^8$ {M$_\odot$}. They also found a lack of correlation between $M$(\hi)\ and $L_\mathrm{B}$ and interpreted it in terms of gas external origin.
Early studies were mostly carried out with pointed single-dish observations aimed at determining global quantities like integrated {\sc H\,i}\ mass and velocity width of ETGs. Recent progress in this kind of analysis has been made possible by single-dish blind surveys of large areas of the sky such as HIPASS (\citealt{2001MNRAS.322..486B}; see \citealt{2002ASPC..273..215S} for an {\sc H\,i}\ study of ETGs based on these data) and ALFALFA \citep{2005AJ....130.2598G}. In particular, the latter pushes the $M$(\hi)\ detection limit below $10^8$ {M$_\odot$}\ for galaxies within a few tens of Mpc from us, allowing a significant increase in the number of detected ETGs. ALFALFA data show that the ETG {\sc H\,i}\ detection rate is a strong function of environment density, from just a few percent inside the Virgo galaxy cluster \citep{2007A&A...474..851D} to about 40 percent outside it (\citealt{2009A&A...498..407G}; see also \citealt{1986A&A...165...15C}). This result fits with the idea that {\sc H\,i}\ is easily stripped from galaxies inside clusters \citep[e.g.,][]{1983AJ.....88..881G}, and with the fact that recent episodes of star formation occur mostly in ETGs in poor environments \citep[e.g.,][]{2010MNRAS.404.1775T}.
While single-dish observations have the advantage of reaching good sensitivity in relatively short integrations, they lack the angular resolution needed to study the detailed {\sc H\,i}\ morphology and kinematics. Starting from the 1980's an increasing number of galaxies was observed at higher angular resolution with interferometers to perform such analysis \citep[e.g.,][]{1980A&A....82..314S,1981ApJ...246..708R,1984A&A...138...77K,1984MNRAS.210..497S,1986AJ.....91..791V,1987ApJ...314...57L,1987A&A...175....4S,1988ApJ...330..684K,1991A&A...243...71V,1994ApJ...423L.101S,1995ApJ...444L..77S}. These and many later observations revealed a surprisingly large diversity of {\sc H\,i}\ morphologies, ranging from very extended (tens of kpc), low column-density discs and rings to unsettled gas distributions indicative of recent gas accretion, gas stripping, or galaxy interaction and merging (see the review by \citealt{1997ASPC..116..310V} and \citealt{2001ASPC..240..657H}). These results revealed {\sc H\,i}\ as a fundamental tracer of the assembly history of ETGs, and prompted work aimed at increasing the size of a sample with deep, homogeneous high-resolution {\sc H\,i}\ data.
A significant step in this direction was made by the combined study of \cite[herafter M06]{2006MNRAS.371..157M} and \cite[herafter O10]{2010MNRAS.409..500O} of 33 galaxies in the SAURON sample. Using data taken with the Westerbork Synthesis Radio Telescope they detected about 10 percent and 2/3 of all galaxies inside and outside the Virgo cluster, respectively (going down to {\sc H\,i}\ masses of a few times $10^6$ {M$_\odot$}). Exploiting the high resolution of their data they show that the {\sc H\,i}\ is distributed on regular discs and rings in about half of the detections (1/3 of all galaxies), with radius ranging from $\sim1$ to tens of kpc \citep[see also][]{2007A&A...465..787O}.
\begin{figure*}
\includegraphics[width=18cm]{figures/a3dhi_fig1.pdf}
\caption{Distribution of $M_\mathrm{K}$ (left), $\Sigma_{10}$ (middle) and $\Sigma_3$ (right) for galaxies in the ATLAS$^{\rm 3D}$\ sample (red line), \atlas\ \hi\ sample (blue line) and {\sc H\,i}\ SAURON sample (grey filled histogram). See the text for the definition of $\Sigma_{10}$ and $\Sigma_3$.}
\label{fig:sample}
\end{figure*}
Their analyses demonstrate that high sensitivity is crucial to detect the faintest signatures of {\sc H\,i}\ accretion in ETGs, and that these are present in most detected galaxies. Furthermore, high angular resolution and the availability of multi-wavelength data make it possible to connect the {\sc H\,i}\ to stars and multi-phase interstellar medium within the host galaxy. This is crucial to understand the role of {\sc H\,i}\ in replenishing ETGs with cold gas, which could then fuel star- and disc formation. For example, M06 and O10 found that all ETGs surrounded by settled {\sc H\,i}\ distributions host ionised gas within $1 R_e$, and that the two gas phases share the same kinematics. Furthermore, all galaxies with {\sc H\,i}\ within $1 R_e$ are detected in CO, and these systems are more likely to be detected in radio continuum too, indicating that some star formation is occurring (see O10 for a discussion of the origin of the radio emission).
The main limitation of M06 and O10 is that the SAURON sample is small, hampering a statistically strong study of the relation between {\sc H\,i}\ and other galaxy properties. For example, the number of galaxies was not sufficient to establish or rule out trends between {\sc H\,i}\ and ETG dynamics or stellar populations, nor to study trends with environment beyond the simple Virgo vs. non-Virgo dichotomy. Also, the uncertainty on the fraction of ETGs hosting regular, rotating {\sc H\,i}\ systems (undoubtedly a very interesting class of objects) was very large.
Another limitation of the SAURON sample is that galaxies were selected to be evenly distributed on the $M_\mathrm{B}$-ellipticity plane rather than to follow the ETG luminosity function, so that the interpretation of the results in terms of galaxy evolution is not straightforward. Works based on the HIPASS and ALFALFA surveys do not suffer from this limitation since both are blind surveys of large regions of the sky. However, they lack the resolution and sensitivity necessary to provide a picture as revealing as that emerging from M06 and O10 (and, in the case of ALFALFA, the sample studied so far is only 40\% larger than the SAURON sample within their common $M_\mathrm{B}$ range).
To overcome these limitations we extend here the study of M06 and O10 to a complete, volume-limited sample of 166 nearby ETGs ($\sim5$ times more galaxies) belonging to the ATLAS$^{\rm 3D}$\ sample, while maintaining approximately the same sensitivity and resolution of their observations. The large size and better selection of the sample studied here allow us to establish the detailed {\sc H\,i}\ properties of ETGs in the local Universe on a firm statistical basis. The availability of a large range of multi-wavelength data taken as part of the ATLAS$^{\rm 3D}$\ project is another element of continuity with work on the SAURON sample, and a major step forward relative to other studies.
The aim of this article is to present the {\sc H\,i}\ properties of galaxies in the sample, how they depend on galaxy luminosity and environment, and their relation to signatures of star formation in the host galaxy. Later work will explore how {\sc H\,i}\ in ETGs relates to other galaxy properties derived from different datasets available within the ATLAS$^{\rm 3D}$\ project, and the connection to simulations. This article is structured as follows. In Sec. \ref{sec:sample} we introduce the galaxy sample. In Sec. \ref{sec:data} we describe radio observations and data reduction. In Sec. \ref{sec:morph} we describe the {\sc H\,i}\ morphology of the detected galaxies and introduce a classification scheme based on it. We also discuss the signatures of star formation in galaxies with different {\sc H\,i}\ morphology, and the relation between {\sc H\,i}\ and H$_2$. In Sec. \ref{sec:detlim} we calculate upper limits on $M$(\hi)\ of undetected galaxies. In Sec. \ref{sec:mf} we discuss the {\sc H\,i}\ mass function of ETGs. In Sec. \ref{sec:spir} we compare the distribution of $M$(\hi), $M$(\hi)/\lk\ and {\sc H\,i}\ column density of ETGs and spirals. In Sec. \ref{sec:env} we investigate the relation between {\sc H\,i}\ properties, galaxy luminosity and environment. Finally in Sec. \ref{sec:summ} we summarise our findings and draw conclusions. \rm
\section{Sample}
\label{sec:sample}
We study the {\sc H\,i}\ properties of galaxies in the volume-limited ATLAS$^{\rm 3D}$\ sample, which includes 260 ETGs within 42 Mpc and brighter than {$M_\mathrm{K}$}$=-21.5$ (Paper I). ETGs are selected from a parent sample as galaxies without spiral arms in available optical images.
Our {\sc H\,i}\ study is based on data taken with the Westerbork Synthesis Radio Telescope (WSRT). For observability reasons we select only the 170 galaxies with declination $\delta \geq 10$ deg. We also exclude all 4 galaxies closer than 15 arcmin to the centre of the Virgo galaxy cluster (NGC~4476, NGC~4478, NGC~4486 and NGC~4486A; observations close to Virgo~A do not provide {\sc H\,i}\ data of sufficient quality). Therefore, we study here a sub-sample of 166 ATLAS$^{\rm 3D}$\ ETGs, 61 percent of the total. We refer to this as the \atlas\ \hi\ sample. Of these galaxies, 39 reside inside the Virgo galaxy cluster and 127 outside it (Paper I). All 166 galaxies are listed in Table \ref{tab:sample} together with properties derived in the present work.
Figure \ref{fig:sample} shows the distribution of galaxy absolute magnitude $M_K$, large-scale environment density $\Sigma_{10}$ and local environment density $\Sigma_{3}$ for the \atlas\ \hi\ sample and the full ATLAS$^{\rm 3D}$\ sample. These parameters are given in Paper I ($M_K$) and in Paper VII ($\Sigma_{10}$ and $\Sigma_{3}$; these are defined as the number density of galaxies contained within a 600-km/s-deep cylinder whose radius is equal to the distance from the tenth and third closest galaxy, respectively). Fig. \ref{fig:sample} shows that the distribution of these parameters is essentially the same for the \atlas\ \hi\ sample and the full ATLAS$^{\rm 3D}$\ sample.
This {\sc H\,i}\ survey expands the study performed by M06 and O10 on the SAURON sample. All but one of the 33 galaxies studied by M06 and O10 belong to the ATLAS$^{\rm 3D}$\ sample. Figure \ref{fig:sample} shows the distribution of galaxies common to the two samples. SAURON galaxies are not distributed very differently if one takes into account statistical uncertainties. The main improvement of the present study are the selection of the sample (which, as discussed in the previous section, is unbiased in galaxy luminosity) and sample size, so that we now have a more representative sampling of the ETG properties. \rm
\section{{\sc H\,i}\ data}
\label{sec:data}
In this section we describe radio observations and data reduction. For some galaxies data were taken as part of earlier studies. Reference to the corresponding papers is given in the text below and in Table \ref{tab:sample}.
\subsection{Observations}
We follow two different observational strategies for galaxies inside and outside the Virgo cluster. We observe all but one of the 127 galaxies outside Virgo with the WSRT (the only exception is NGC~4762, for which we use ALFALFA data). WSRT data were taken as part of previous projects for 18 of these 127 objects \citep[M06 and O10]{2009A&A...494..489J}. Another 3 fall within the field of view of these earlier observations. All remaining 105 galaxies are observed for ATLAS$^{\rm 3D}$. Both old and new data are taken using the WSRT in the maxi-short configuration. We observe using a bandwidth of 20 MHz sampled with 1024 channels, corresponding to a recessional velocity range of $\sim4000$ km/s and a channel width of $\sim4$ km/s for the {\sc H\,i}\ emission line. The only exception is NGC~2685, observed by \cite{2009A&A...494..489J} with 1024 channels over a 10-MHz bandwidth.
We observe each galaxy for 12 h. For some galaxies, earlier data are deeper than our 12 h integration (see Table \ref{tab:sample}, column 8) and we use those in our analysis.
Data are less homogeneous for the 39 galaxies inside Virgo. We take WSRT data with the same array configuration and correlator set-up described above for 3 galaxies. Equivalent data are available for 13 more galaxies in O10. A field including 2 more ATLAS$^{\rm 3D}$\ galaxies was observed with the WSRT by Oosterloo and collaborators with the same WSRT configuration and correlator set-up described above, and we use their data in our analysis. We use the ALFALFA database to look for {\sc H\,i}\ emission in the remaining 21 Virgo galaxies\footnote{\url{http://arecibo.tc.cornell.edu/hiarchive/alfalfa}.}. We find 2 galaxies with a possibly associated ALFALFA {\sc H\,i}\ detection: NGC~4694 and NGC~4710. We make use of VLA data taken for the VIVA project to study the former \citep{2009AJ....138.1741C}; and we observe the latter with the WSRT using the same array configuration and correlator set-up as above. We use ALFALFA spectra to derive $M$(\hi)\ upper limits for the remaining 19 galaxies (see Sec. \ref{sec:detlim}).
\subsection{WSRT data reduction and products}
We reduce the WSRT data in a standard way using a dedicated pipeline based on the Miriad reduction package \citep{1995ASPC...77..433S}. The standard pipeline products are {\sc H\,i}\ cubes built using robust=0 weighting and 30-arcsec-FWHM tapering (see \citealt{1995AAS...18711202B} for an explanation of the robust parameter). The discussion below is based on the analysis of these cubes unless otherwise stated. The pipeline is validated by comparing its products to those obtained with manual reduction by M06 and O10 for galaxies in common.
The typical angular resolution of the cubes is $35\times35/\sin\left(\delta\right)$ arcsec$^2$, where $\delta$ is the declination. All cubes are made at a velocity resolution of 16 km/s (after Hanning smoothing). The noise ranges between 0.4 and 0.7 mJy/beam. The median noise is 0.5 mJy/beam, and 90 percent of the cubes have noise below 0.6 mJy/beam. At the median noise level the 5$\sigma$ column density threshold within one velocity resolution element is $\sim3.6\times10^{19}\times \sin\left(\delta\right)$ cm$^{-2}$. The \atlas\ \hi\ sample covers the $\delta$ range 10 - 60 deg, corresponding to $N$(\hi)\ in the range 0.6 - 3.1 $\times10^{19}$ cm$^{-2}$.
For each galaxy we build a total {\sc H\,i}\ image by summing along the velocity axis all emission included in a mask constructed as follows. A pixel belongs to the mask if the absolute value of its flux density is above 4$\sigma$ in at least one of the following cubes: \it (i) \rm the original cube; \it (ii) \rm the three cubes obtained smoothing the original cube in velocity with a Hanning filter of FWHM 32, 64 and 112 km/s; \it (iii) \rm the original cube smoothed with a 60-arcsec-FWHM Gaussian beam; \it (iv) \rm the three cubes obtained by Hanning smoothing the 60-arcsec cube as in point \it (ii)\rm. This allows us to be sensitive to {\sc H\,i}\ emission over a wide range of angular and velocity scales. The binary masks are then enlarged by convolution with the synthesised beam, and only mask pixels whose value is above 50 percent of the convolved mask peak are kept. This removes a large fraction of spurious detections with small angular size. It also ensures that we do not miss faint gas just below the detection threshold at the edge of the {\sc H\,i}\ distribution.
Total {\sc H\,i}\ images are obtained as the zero-th moment of the masks built following the above procedure. These images contain also pixels with negative surface brightness because we apply the detection algorithm to the absolute value of the flux density. We use these ``negative'' pixels to estimate the column density detection threshold of the images. In practice, we determine the surface brightness threshold $-S_\mathrm{th}$ below which only 5 percent of the negative pixels are retained. The final total {\sc H\,i}\ image includes only pixels with $|S|\geq +S_\mathrm{th}$. The typical value of $S_\mathrm{th}$ corresponds to an {\sc H\,i}\ column density of a few times $10^{19}$ cm$^{-2}$, the exact value varying from image to image.
We also build {\sc H\,i}\ cubes with robust=1 weighting (close to natural weighting) and no tapering. Below we refer to this weighting scheme as $R01$. The noise in $R01$ cubes is a factor of $\sim1.5$ lower than in the standard cubes. However, its pattern is very patchy and the overall image quality is lower. Therefore, we use these cubes only when the detection algorithm described above reveals low-surface-brightness emission missing from the standard cubes (see Table \ref{tab:sample}, column 7). In these cases, the {\sc H\,i}\ mass given in Table \ref{tab:sample} is derived from the $R01$ cube.
We note that the $R01$ cubes have slightly better angular resolution than the standard cubes: $25\times25/sin(\delta)$ arcsec$^2$. Therefore, the formal column density sensitivity is $\sim30$ percent worse.
\section{{\sc H\,i}\ morphology and its relation to the host galaxy}
\label{sec:morph}
We detect 53/166 = 32 percent of all ETGs in the \atlas\ \hi\ sample. Unlike the detection rate presented in M06 and O10, this does not include a number of cases where {\sc H\,i}\ is detected only in absorption or where we cannot securely assign {\sc H\,i}\ to the observed ETG because of confusion with neighbouring galaxies (see below).
The {\sc H\,i}\ detection rate depends strongly on environment density, being 4/39=10 percent inside Virgo and 49/127=39 percent outside it (see Sec. \ref{sec:env} for a comparison to previous results). We postpone a full discussion of environmental effects to Sec. \ref{sec:env}. In this section we describe the {\sc H\,i}\ morphology of all detected galaxies regardless of their environment.
\subsection{{\sc H\,i}\ morphological classes}
We show total {\sc H\,i}\ images of all detected ETGs in Fig. \ref{fig:gallery}. The figure suggests the existence of a few broad types of {\sc H\,i}\ morphology, which we describe with the following {\sc H\,i}\ morphological classes:
\begin{itemize}
\item $D$ (large discs) -- most of the {\sc H\,i}\ rotates regularly and is distributed on a disc or ring larger than the stellar body of the galaxy.
\item $d$ (small discs) -- most of the {\sc H\,i}\ rotates regularly and is distributed on a small disc confined within the stellar body of the galaxy.
\item $u$ (unsettled) -- most of the {\sc H\,i}\ exhibits unsettled morphology (e.g., tails or streams of gas) and kinematics.
\item $c$ (clouds) -- the {\sc H\,i}\ is found in small, scattered clouds around the host galaxy.
\end{itemize}
\noindent After examining morphology and kinematics of the detected {\sc H\,i}\ we assign each galaxy to one of these classes. The {\sc H\,i}\ class of detected objects is given in Table \ref{tab:sample} and indicated in Fig. \ref{fig:gallery}. Further notes on the {\sc H\,i}\ morphology and kinematics of individual galaxies are given in Table \ref{tab:sample}, column 7.
\begin{table}
\centering
\caption{Number of galaxies in the various {\sc H\,i}\ classes as a function of Virgo membership (see Sec. \ref{sec:env}) and the presence of CO and dust/blue features (see Sec. \ref{sec:h2}).}
\begin{tabular}{rrrrrr}
\hline
{\sc H\,i}\ & all & inside & outside & CO & dust/blue\\
class & galaxies & Virgo & Virgo &det. & features\\
\hline
all & 166 &39 &127 & 38 & 36 \\
undet. & 113 & 35 &78 & 15 & 13\\
det. & 53 & 4 & 49 & 23 & 23 \\
\\
$D$ & 24 & 1 & 23 & 10 & 10 \\
& (10) & (1) & (9) & (1) & (2) \\
$d$ & 10 & 1 & 9 & 8 & 10 \\
$u$ & 14 & 2 & 12 & 5 & 3 \\
& (4) & (0) & (4) & (0) & (0) \\
$c$ & 5 & 0 & 5 & 0 & 0\\
\hline
\end{tabular}
Numbers in parenthesis indicate the number of galaxies within that class for which {\sc H\,i}\ is not found in the centre of the host galaxy. All $d$'s and none of the $c$'s host {\sc H\,i}\ in their centre.
\label{tab:summary}
\end{table}
This classification scheme is an attempt to simplify the variety of {\sc H\,i}\ morphologies seen in Fig. \ref{fig:gallery}. It also reflects our view that different {\sc H\,i}\ morphologies give different indications on the assembly and gas-accretion history of the host galaxy. For example, $D$'s and $d$'s are galaxies with a relatively quiet recent history on a time-scale proportional to the orbital time of the rotating gas (this is obviously larger in $D$ than $d$ systems; as an example, it is $\sim1$ Gyr for the $D$ galaxy NGC~6798, and $\sim200$ Myr for the $d$ NGC~5422). On the contrary, galaxies in class $u$, where most of the gas is not rotating regularly around the stellar body, have recently experienced tidal forces or episodes of gas accretion/stripping involving a large fraction of their gas reservoir. In other words, their {\sc H\,i}\ content is evolving, and for some of them the current {\sc H\,i}\ properties may be just a transient phase. In contrast, in $D$ objects the {\sc H\,i}\ has likely been part of the galaxy (and orbiting around the stellar body) for many gigayears. Finally, galaxies in class $c$ are surrounded by {\sc H\,i}\ not obviously associated to them, and we regard them as objects with similar {\sc H\,i}\ properties as undetected ETGs.
\begin{figure*}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC3945.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC5103.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC1023.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC0680.pdf}
\includegraphics[width=43mm]{figures/NGC3945_testpv.pdf}
\includegraphics[width=43mm]{figures/NGC5103_testpv.pdf}
\includegraphics[width=43mm]{figures/NGC1023_testpv.pdf}
\includegraphics[width=43mm]{figures/NGC0680_testpv.pdf}
\caption{A sequence of {\sc H\,i}-rich ETGs with increasingly less regular gas configurations (left to right). The sequence shows galaxies with very large {\sc H\,i}\ distributions relative to the stellar body. The top row shows total-{\sc H\,i}\ contour images. We refer to the caption of Fig. \ref{fig:gallery} for a description of the content of each image. In this figure the top-right scale bar indicates 10 kpc at the galaxy distance. The bottom row shows position-velocity diagrams of galaxies in the top row drawn along an axis which highlights relevant features in the {\sc H\,i}\ kinematics (see text). We plot angular and velocity offset along horizontal and vertical axis, respectively. The diagrams are drawn along an axis whose position angle is indicated on the bottom right of each diagram (north to east). Contours are drawn at $1.0\times2^n$ mJy beam$^{-1}$, $n=0,1,2,...$. The cross on the bottom left indicates 10 kpc along the horizontal axis and 50 km s$^{-1}$ along the vertical axis.}
\label{fig:sequenced}
\end{figure*}
Similarly, the distinction between $D$ and $d$ galaxies, which is entirely based on the size of the {\sc H\,i}\ distribution, is motivated by our view that they are fundamentally different objects (see Sec. \ref{sec:4.3}). In order to make this classification as objective as possible we define the {\sc H\,i}\ radius $R$(\hi)\ as the maximum distance of the $N$(\hi)=$5\times10^{19}$ cm$^{-2}$ isophote from the galaxy centre (deconvolved with the beam size). We then compare $R$(\hi)\ to the optical effective radius $R_e$ given in Paper I. We adopt $R$(\hi)$=3.5\times R_e$ as the dividing line between small and large {\sc H\,i}\ systems (for a de Vaucouleur profile this radius includes 90 percent of the total light).\footnote{The only galaxy whose classification would not fit in our picture is NGC~3414, which we regard as hosting a large {\sc H\,i}\ system although its $R$(\hi)\ is smaller than 3.5$\times R_e$. However, the gas in this galaxy is distributed on a polar configuration, so that it may be more appropriate to compare $R$(\hi)\ to $R_e\times(1-\epsilon)$ where $\epsilon$ is the optical ellipticity given in Paper II. When this is done the ratio between {\sc H\,i}\ and optical radius is larger than 3.5 so that we classify this galaxy as $D$.}
Given this classification scheme, $D$ galaxies have $M$(\hi)\ in the range $10^8$-$10^{10}$ {M$_\odot$}, and the gas extends out to many tens of kpc from the centre of the galaxy (in some cases more than 10 $R_\mathrm{e}$). In contrast, $M$(\hi)\ in $d$ systems is typically below $10^8$ {M$_\odot$}, and the size of the gas distribution is up to a few kpc. Galaxies in group $u$ contain between a few times 10$^7$ and 10$^{10}$ {M$_\odot$}\ of {\sc H\,i}, often stretching to many tens of kpc from the galaxy. Finally, galaxies in group $c$ have $M$(\hi)\ below a few times 10$^7$ {M$_\odot$}, comparable to the {\sc H\,i}\ mass upper limits described in Sec. \ref{sec:detlim}.
\subsection{A continuum of {\sc H\,i}\ morphologies}
\label{sec:4.2}
We find 24, 10, 14, and 5 galaxies in group $D$, $d$, $u$ and $c$, respectively, corresponding to 14, 6, 8 and 3 percent of the sample. Only 1 $D$, 1 $d$ and 2 $u$'s belong to the Virgo cluster. Outside Virgo 18, 7, 9 and 4 percent of all galaxies belong to group $D$, $d$, $u$ and $c$, respectively. About 40 percent ($10/24$) of all $D$'s are rings (or discs with a central {\sc H\,i}\ hole). We summarise these results in Table \ref{tab:summary}.
The above figures show that within the ATLAS$^{\rm 3D}$\ volume as many as 1/5 of all ETGs (and 1/4 of all ETGs outside Virgo) host rotating {\sc H\,i}\ distributions (groups $D$ and $d$). These objects represent the majority of our detections, 64 percent, in agreement with M06 and O10. Therefore, \it if {\sc H\,i}\ is detected in an ETG it will most likely be distributed on a disc or a ring. \rm
Within the adopted classification scheme it is not always easy to assign a galaxy to a given class. In fact, we find a number of objects intermediate (or in transition) between classes. For the sake of simplicity we will make use of this classification in the rest of the article, but the \atlas\ \hi\ sample reveals rather a continuum of {\sc H\,i}\ morphologies going from settled, rotating systems to progressively more and more disturbed ones, and from the latter to systems of scattered {\sc H\,i}\ clouds which may or may not be related to the nearby ETG. We guide the reader through this continuum of morphologies in the rest of this section.
Figure \ref{fig:sequenced} shows a sequence of increasingly less regular {\sc H\,i}\ distributions (left to right) whose size is large relative to that of the host galaxy. The sequence shows the continuity of {\sc H\,i}\ morphology going from the most regular $D$ objects (left) to the most irregular, large $u$'s (right). It begins on the left with an {\sc H\,i}\ ring, likely a resonance related to a bar (see also the stellar ring coincident with the {\sc H\,i}\ distribution). The morphology (top row) and kinematics (bottom row) of the {\sc H\,i}\ ring are extremely regular. The sequence continues on the second panel with a disc whose rotation and morphology are somewhat less regular. In this galaxy a tail of gas is visible to the west, and a small neighbour to the east might be interacting with our target. Although the {\sc H\,i}\ kinematics exhibits an overall ordered rotation (but misaligned relative to the stellar kinematics), it is clear that some of the gas has not settled yet within the galaxy potential (or has been disturbed recently). Systems like this provide a link between $D$'s and galaxies in group $u$, where most (but not necessarily all) of the gas is unsettled.
The third panel shows a galaxy that we classify as $u$. However, there are hints that at least part of the {\sc H\,i}\ is rotating around the stellar body (see the position-velocity diagram on the bottom row and the description of this galaxy in \citealt{1984MNRAS.210..497S} and M06). Systems like this could move towards the $D$ class if their evolution proceeds without significant merger/accretion events for a few gas orbital periods. Finally, the fourth panel shows completely unsettled {\sc H\,i}\ (\citealt{2011arXiv1105.5654D}, herafter Paper IX, discuss the optical signature of a recent merger in this galaxy). It is possible to see the sequence in Fig. \ref{fig:sequenced} as one of decreasing time passed since the last major episode of gas accretion or stripping (relative to the gas orbital period).
A similar sequence can be drawn for {\sc H\,i}-detected galaxies where the size of the gas distribution is comparable to that of the stellar body. Such a sequence, shown in Fig. \ref{fig:sequencee}, illustrates the continuity of {\sc H\,i}\ morphology going from the most regular $d$ objects (left) to the most irregular $u$'s (right). The sequence starts with a very regular small {\sc H\,i}\ disc. It continues with a very faint, small disc connected spatially and in velocity to a cloud outside the stellar body. This system appears fairly regular within the stellar body (given the low signal-to-noise ratio), but the outer cloud indicates that some accretion may be on-going. In fact, the {\sc H\,i}\ cloud suggests that small {\sc H\,i}\ discs may form by accreting gas from small companions or the surrounding medium (O10).
The third panel in Fig. \ref{fig:sequencee} shows a system which is difficult to classify. We detect {\sc H\,i}\ only on one side of the galaxy, and because of this we classify it as $u$. However, the position-velocity diagram shows a hint that gas may be present also on the other side of the galaxy at opposite velocity relative to systemic. This is the behaviour expected for a disc, so this galaxy may actually be a misclassified $d$. We conclude the sequence with a system where {\sc H\,i}\ is detected far from the stellar body but is clearly connected to it. The modest amount of gas and the size of the {\sc H\,i}\ distribution relative to the stellar body suggests that this system may evolve towards the $d$ class if left undisturbed.
\begin{figure*}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC5422.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC3489.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC7280.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC2768.pdf}
\includegraphics[width=43mm]{figures/NGC5422_testpv_sequence.pdf}
\includegraphics[width=43mm]{figures/NGC3489_testpv_sequence.pdf}
\includegraphics[width=43mm]{figures/NGC7280_testpv_sequence.pdf}
\includegraphics[width=43mm]{figures/NGC2768_testpv_sequence.pdf}
\caption{A sequence of {\sc H\,i}-rich ETGs with increasingly less regular gas configurations (left to right). The sequence shows galaxies with {\sc H\,i}\ distributions of size similar to that of the stellar body. See the caption of Fig. \ref{fig:sample} for a description of the images. In this figure, position-velocity diagrams are drawn using the $R01$ cubes. Contours are drawn at $N_0\times2^n$ mJy beam$^{-1}$, $n=0,1,2,...$, where $N_0=0.5$ for NGC~2768, 0.75 for NGC~5422 and NGC~3489, 1.0 for NGC~7280.}
\label{fig:sequencee}
\end{figure*}
\begin{figure*}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC4111.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC5557.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC3193.pdf}
\includegraphics[width=43mm]{figures/HISPECIALfigures/NGC4521.pdf}
\caption{A sequence of {\sc H\,i}-rich ETGs illustrating the continuity between class $u$ and class $c$. We refer to the caption of Fig. \ref{fig:gallery} for a description of the content of each image. In this figure the top-right scale bar indicates 10 kpc at the galaxy distance.}
\label{fig:sequenceb}
\end{figure*}
Another step in this continuum of {\sc H\,i}\ morphologies is illustrated in Fig. \ref{fig:sequenceb}. In this figure we show that some the unsettled systems appear as a bridge between class $u$ and class $c$. The sequence starts from an extended, unsettled {\sc H\,i}\ system. The large number of gas-rich neighbours suggests that galaxy interaction may be responsible of the disturbed {\sc H\,i}\ morphology \citep[see also][]{2001ASPC..240..867V}. The second and third galaxies on the sequence are surrounded by a few {\sc H\,i}\ clouds forming a coherent gaseous system -- in both cases the clouds are connected to each other and to the ETG in velocity and are likely to be the densest clumps of an {\sc H\,i}\ tail (for NGC~5557 see Paper IX). These systems represent a natural link to galaxies in group $c$ (represented by the fourth object in the sequence), where scattered clouds are detected around an ETG but their connection to it is much less obvious.
The final piece in this continuum of morphologies is provided by the continuity between $c$ and undetected objects. Many $c$ galaxies are difficult to classify as it is sometimes not clear whether we are seeing the densest gas clumps of a $u$ system, or gas is instead floating unbound in the inter-galactic medium. In fact, a number of ETGs which we classify as non detections are surrounded by {\sc H\,i}\ with properties similar to those of $c$ galaxies. We show these objects in Fig. \ref{fig:gallery_undet_a}, (the figure includes also galaxies close to or interacting with an {\sc H\,i}-rich neighbour, so that confusion prevents us from establishing whether they contain any {\sc H\,i}). This ambiguity has little effect on our discussion in the following sections as the {\sc H\,i}\ mass of $c$ objects is comparable to the detection limit of our observations (see Sec. \ref{sec:detlim}).
\subsection{Small and large {\sc H\,i}\ discs: relation to the kinematics of the host galaxy}
\label{sec:4.3}
At the beginning of this section we introduced a criterion to distinguish between large and small rotating {\sc H\,i}\ systems (i.e., $D$ and $d$). We consider this distinction important as there are fundamental differences between these objects (besides their size). As mentioned, $D$'s contain between $10^8$ and $10^{10}$ {M$_\odot$}\ of {\sc H\,i}, while in $d$'s $M$(\hi)\ is typically below $10^8$ {M$_\odot$}. Furthermore, the {\sc H\,i}\ kinematics is aligned with the stellar kinematics in most galaxies belonging to group $d$ -- 8/10 objects. Figure \ref{fig:pv} shows position-velocity diagrams of all $d$'s along the stellar kinematical major axis given in Paper II (perpendicular to it in the case of NGC~3499 -- see below). Comparing these diagrams with the stellar velocity fields published in Paper II we see that the {\sc H\,i}\ rotation is aligned with that of the stars in all galaxies except NGC~3032, where the {\sc H\,i}\ is counter-rotating, and NGC~3499, where the {\sc H\,i}\ is polar. Even in these two cases a connection between {\sc H\,i}\ and stellar body exists. In the former the gas disc is aligned with a kinematically-decoupled stellar component \citep[O10]{2006MNRAS.373..906M}. In the latter the {\sc H\,i}\ is aligned with a faint stellar disc and a dust lane visible in the SDSS image.
\begin{figure*}
\includegraphics[width=33mm]{figures/NGC2824_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC3032_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC3182_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC3489_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC3499_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC4150_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC4710_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC5422_testpv.pdf}
\includegraphics[width=33mm]{figures/NGC5866_testpv.pdf}
\includegraphics[width=33mm]{figures/UGC05408_testpv.pdf}
\caption{Position-velocity diagrams of all $d$ galaxies along the stellar kinematical major axis given in Paper II (perpendicular to it for NGC~3499). In all diagrams the x axis represents angular offset and the y axis velocity offset. Dotted lines indicate the centre of the galaxy. The diagrams are drawn using the $R01$ cubes for all galaxies except NGC~3032. Contours start at 1.0 mJy beam$^{-1}$ for NGC~3032, at 0.6 mJy beam$^{-1}$ for NGC~3182, NGC~3499 and NGC~4150, and at 0.75 mJy beam$^{-1}$ for all other galaxies. Contours increase by a factor of two at each step. The cross on the bottom left indicates 2 kpc along the horizontal axis and 50 km s$^{-1}$ along the vertical axis.}
\label{fig:pv}
\end{figure*}
In contrast with this result, visual inspection of the stellar velocity fields (Paper II) and the HI cubes shows that misaligned discs/rings are found in more than half (14/24) of all $D$ galaxies. This includes a number of polar rings and counter-rotating discs/rings (see Fig. \ref{fig:gallery} and column 7 in Table \ref{tab:sample}). From a kinematical point of view, it seems that $d$ {\sc H\,i}\ distributions are tightly coupled to their host galaxy, while in $D$'s the {\sc H\,i}\ kinematics is often different from the stellar one. A future paper will address the relation between {\sc H\,i}\ and stellar kinematics in more details, including the classification of ETGs in the ATLAS$^{\rm 3D}$\ sample in slow- and fast rotators (Paper III). \rm
\subsection{{\sc H\,i}, H$_2$ and star formation signatures}
\label{sec:h2}
As part of their study of 33 ETGs belonging to the SAURON sample, O10 show that all galaxies where {\sc H\,i}\ is detected within $\sim1R_\mathrm{e}$ contain molecular gas in their centre, and that H$_2$ dominates the ISM in this region. On the contrary, the CO detection rate is $\sim20$ percent for ETGs with no central {\sc H\,i}\ detection. Here we revisit their result by comparing the {\sc H\,i}\ properties of ETGs in the \atlas\ \hi\ sample to results of Paper II (which provides a catalogue of dust discs, dusty filaments and blue regions in all ETGs in the ATLAS$^{\rm 3D}$\ sample) and Paper IV (where we discuss the ATLAS$^{\rm 3D}$\ CO survey).
We detect dust/blue features and CO in 36 and 38 of the 166 ETGs in the \atlas\ \hi\ sample, respectively. The presence of dust/blue features and molecular gas are very tightly related to one another (Paper IV), so that most galaxies with CO also have dust/blue features and vice-versa. With reference to Table \ref{tab:summary}, dust/blue features and CO are present in $43\pm9$ percent of all {\sc H\,i}\ detections. In contrast the dust/blue and CO detection rate is $13\pm3$ percent for the {\sc H\,i}\ non-detections. Therefore, galaxies detected in {\sc H\,i}\ are $\sim3$ times more likely to contain signatures of star formation than those with no detectable neutral hydrogen.
We make use of the classification presented above to investigate how {\sc H\,i}\ morphology is related to the occurrence of star formation. In the previous section we discuss the kinematical link between {\sc H\,i}\ and stellar body in $d$ ETGs. This result is complemented by the detection of star formation signatures in all these galaxies. As illustrated in Table \ref{tab:summary}, dust/blue features are detected in $10/10=100\pm32$ percent of them\footnote{Note that NGC~4150 is not listed as containing dust/blue features in Paper II because of the poor SDSS image quality. However HST observations by \cite{2011ApJ...727..115C} show that also this galaxy is characterised by a dusty and star-forming core.}. Furthermore, we detect CO in $8/10=80\pm28$ percent of them. The exceptions are NGC~3499 and NGC~5422, for which the upper limit on $M$(H$_2$)\ is $\sim5\times10^7$ {M$_\odot$}.
A further indication that stars are being formed efficiently in $d$ galaxies comes from the analysis of the ISM composition in their central region. As done in O10, we compare the H$_2$ mass given in Paper IV (measured within the $\sim$20 arcsec IRAM beam) to the {\sc H\,i}\ mass contained in one WSRT beam (whose minor axis is $\sim30$ arcsec) centred on the galaxy. This comparison shows that in $d$'s the $M$(H$_2$)/$M$(\hi)\ ratio is high -- always above $\sim1$, and $>10$ in 5 out of 8 galaxies detected in CO. Therefore, the ISM in these galaxies has similar properties as that in the central region of spirals \citep{2008AJ....136.2782L}, confirming that in many of them {\sc H\,i}\ is being turned into H$_2$ efficiently.
Signatures of star formation are less common in $D$ galaxies and depend on whether the {\sc H\,i}\ is distributed on a ring (10 objects -- see Table \ref{tab:summary}) or extends all the way to the centre of the galaxy (14 objects). We detect CO in 9 of the ETGs with central {\sc H\,i}\ ($64\pm21$ percent) and dust/blue features in 8 of them ($57\pm20$ percent). In contrast only 1 galaxy with an {\sc H\,i}\ ring is detected in CO ($10\pm10$ percent) and 2 (including the CO detection) exhibit dust/blue features ($20\pm14$ percent).
The physical state of the ISM in these objects may be different from that in $d$'s as the $M$(H$_2$)/$M$(\hi)\ ratio is somewhat lower. When CO is detected, we find $M$(H$_2$)/$M$(\hi)\ values in the range 0.5 - 3 for 80 percent of the $D$'s, and never larger than $\sim10$. When CO is not detected the upper limit on $M$(H$_2$)/$M$(\hi)\ is always above $\sim1$, so that these systems are not necessarily different from those with a CO detection (they may simply contain too little central {\sc H\,i}\ for molecular gas to be detected). Therefore, also in $D$'s most of the central ISM is found in the form of molecular gas, but the conversion of {\sc H\,i}\ to H$_2$ may be less efficient than in $d$'s. One possibile explanation is that {\sc H\,i}\ in the centre of these galaxies has lower column density.
A similar result holds for $u$ galaxies. These systems are unsettled so that it is not always easy to judge whether {\sc H\,i}\ is present within the stellar body, or is simply projected onto it on the sky. However, {\sc H\,i}\ is certainly not present in the central region of NGC~3193, NGC~4026 (a ring with a large and massive {\sc H\,i}\ tail), NGC~5198 and NGC~5557. None of these systems hosts CO or dust/blue features. Of the remaining 10 $u$ galaxies, 3 host dust/blue features ($30\pm17$ percent) and 5 host molecular gas ($50\pm22$ percent). As in $D$'s, the typical (upper limit on) $M$(H$_2$)/$M$(\hi)\ is of the order of unity, and none of these objects has $M$(H$_2$)/$M$(\hi)$>10$.
\begin{figure*}
\includegraphics[width=18cm]{figures/detlim_final.pdf}
\caption{Peak signal to noise $\mathrm{\left(S/N\right)_{max}}$\ of all galaxies detected in the standard, 12-h {\sc H\,i}\ cubes (see text) plotted against the size of the {\sc H\,i}\ system in units of the beam (left), and the spectral line width at the position of $\mathrm{\left(S/N\right)_{max}}$\ (right). The top $H$ symbols represent the $FWHM$ of the Hanning filters used to smooth the {\sc H\,i}\ cubes when searching for emission (see Sec. \ref{sec:data}).}
\label{fig:detlim}
\end{figure*}
Consistent with the above discussion, none of the 5 $c$'s exhibits dust/blue features or is detected in CO. We therefore confirm O10 result that the detection rate of star-formation signatures is high ($\sim2/3$) for ETGs hosting neutral hydrogen within $\sim1R_\mathrm{e}$, so that {\sc H\,i}\ seems to provide material for star formation. In particular, 100 percent of all $d$'s, $\sim60$ percent of all $D$'s with central {\sc H\,i}, and $\sim50$ percent of all $u$'s with central {\sc H\,i}\ exhibit such signatures. At the sensitivity of our data, even galaxies with central {\sc H\,i}\ but no star formation signatures are consistent with the picture that the central ISM of ETGs is dominated by molecular gas -- they may simply contain too little {\sc H\,i}\ for CO to be detected. In contrast with these results, galaxies surrounded by an {\sc H\,i}\ distribution with a central hole, and galaxies with no detectable {\sc H\,i}, show signs of recent star formation in a minority of the cases ($\sim15$ percent).
These trends indicate that {\sc H\,i}\ may play an important role in fuelling rejuvenation and the growth of a stellar disc in ETGs. In particular, in a number of $d$'s (the ETGs with the highest {\sc H\,i}-to-H$_2$ conversion rate in their centre) the {\sc H\,i}\ morphology suggests that recent accretion (possibly from small companions) is the source of the ISM (e.g., NGC~3032, NGC~3489, NGC~4150 -- see O10). Yet, 15/38 CO detections and 13/36 ETGs with dust blue/features are not detected in {\sc H\,i}. The H$_2$ mass of these systems is not particularly low, so they are galaxies with a very high $M$(H$_2$)/$M$(\hi)\ ratio (lower limits on this are above $\sim5$ in most of these cases). Hosting {\sc H\,i}\ at the present time (and at the sensitivity of our observations) is therefore not a necessary condition for an ETG to host star formation. It is possible that some ETGs have been stripped of most their {\sc H\,i}\ but not their molecular gas if they live in a dense environment with a hot medium (see Sec. \ref{sec:env} and Paper IV). However, more than half of the galaxies with signs of star formation but no detected {\sc H\,i}\ live outside Virgo, so this cannot be the only explanation. These results highlight the complex gas accretion history and interplay between different ISM phases of ETGs.
\subsection{{\sc H\,i}\ in absorption}
We conclude this section by noting that {\sc H\,i}\ gas can also be detected in absorption against a radio continuum central source (1.4 GHz continuum images are produced as part of the pipeline described in Sec. \ref{sec:data}).
For each object, we obtain continuum images using the line-free channels. The typical noise of the images is $\sim60$ $\mu$Jy beam$^{-1}$ and more than 40 percent of the galaxies were detected with radio power in the range $10^{18}$-10$^{22}$ W Hz$^{-1}$, consistent with previous studies of the continuum emission of this kind of objects \citep[e.g.,][]{1989MNRAS.240..591S,2007MNRAS.375..931M}. A complete discussion of these results will be presented in a future paper.
We look for {\sc H\,i}\ absorption in all objects with a continuum flux above 8 mJy. This results in three detections of objects where {\sc H\,i}\ is observed only in absorption (NGC~5322, NGC~5353 and PGC~029321), while in three more objects absorption is detected in addition to emission (NGC~2824, NGC~3998 and NGC~5866; see Table \ref{tab:sample}). The continuum flux density of objects detected in {\sc H\,i}\ absorption ranges between 8 and $\sim30$ mJy. The {\sc H\,i}\ absorption is relatively narrow, 50 - 80 km/s, and in all cases it is centred on the systemic velocity. The absorbed flux ranges between 3 and 4 mJy. We derive optical depths of 7 - 10 percent and column densities of $4 - 7 \times 10^{20}$ cm$^{-2}$ (assuming T$_{\rm spin}$ = 100 K).
This absorption may be the result of small disc structures (i.e. discs that are not large enough to be seen in emission if a relatively strong continuum source is present). For example, two of the three galaxies detected only in absorption exhibit dust or blue features within the stellar body, consistent with the result described above for $d$ galaxies. Only one of them is detected in CO (PGC~029321).
It is worth noting that only 19 objects among the $\sim140$ observed with the WSRT have radio continuum flux density $S_{1.4GHz} > 8$ mJy. The detection rate of {\sc H\,i}\ in absorption is therefore $\sim30$ percent for galaxies with sufficient continuum emission. This underlines that the {\sc H\,i}\ detection rate of our study (in emission and absorption) is just a lower limit to the fraction of ETGs containing neutral hydrogen gas.
\section{$M$(\hi)\ detection limits}
\label{sec:detlim}
The analysis presented in the following sections is based on the images in Fig. \ref{fig:gallery}, the classification discussed above, the {\sc H\,i}\ mass measured from these images and listed in Table \ref{tab:sample}, as well as on $M$(\hi)\ upper limits for the undetected galaxies. In this section we discuss how we calculate these upper limits.
The data cubes described in Sec. \ref{sec:data} give the {\sc H\,i}\ surface brightness distribution projected on the sky per unit velocity interval along the line of sight. The detection of {\sc H\,i}\ in a galaxy is therefore driven by its \it peak surface brightness \rm relative to the noise level at the angular and velocity resolution of the data, and not by the total mass of {\sc H\,i}\ hosted by the galaxy. This makes the calculation of upper limits \mhi$_\mathrm{lim}$\ for undetected galaxies rather delicate.
The typical approach in the literature is to calculate \mhi$_\mathrm{lim}$\ as the minimum detectable {\sc H\,i}\ mass within one angular-resolution element centred on the galaxy, and integrating over a velocity interval representative of the gas line-width. For example, if we calculate \mhi$_\mathrm{lim}$\ as a 3$\sigma$ signal within one synthesised beam and within a 200 km/s velocity interval, we obtain \mhi$_\mathrm{lim}$$=2.0 \cdot (d/\mathrm{10\ Mpc})^2 \times 10^6 $ {M$_\odot$}. However, a larger {\sc H\,i}\ mass can be hidden below the noise if it is spread over many beams. The first question we wish to answer is therefore whether we are missing a population of extended, massive {\sc H\,i}\ distributions with column density below the noise of our observations.
To find an answer we calculate the maximum signal-to-noise ratio $\mathrm{\left(S/N\right)_{max}}$\ of the {\sc H\,i}\ emission for galaxies detected in the standard cubes\footnote{We therefore leave out NGC~3499, NGC~4150 and NGC~7332, which are detected only in the $R01$ cubes, and NGC~4406, for which only a total {\sc H\,i}\ image (and not the cube, necessary for this analysis) is available. We note that for some galaxies the available cubes are obtained after multiple 12-h integrations (see Table \ref{tab:sample}). For consistency, we re-analyse a single 12-h integration for these galaxies, and use the result in this section.}. This calculation is made including data-cubes at all angular and velocity resolutions listed in Sec 3 (i.e., all possible combinations of the velocity resolutions 16, 32, 64 and 112 km/s, and of the angular resolutions $\sim30$ and 60 arcsec). The value of $\mathrm{\left(S/N\right)_{max}}$\ indicates how easily a galaxy is detected above the noise.
The left panel in Fig. \ref{fig:detlim} shows $\mathrm{\left(S/N\right)_{max}}$\ plotted versus the area $A$({\sc H\,i}) occupied by the detected gas in units of the beam area $A$(beam). This plot shows a trend of increasing $\mathrm{\left(S/N\right)_{max}}$\ with increasing $A$({\sc H\,i})/$A$(beam). In particular, the majority of the detections have very high $\mathrm{\left(S/N\right)_{max}}$\ (above 10), and this is true for nearly all detections larger than 10 beams. There is no hint of a population of very extended {\sc H\,i}\ systems close to the noise of our observations.
\begin{figure}
\includegraphics[width=8.5cm]{figures/detlim_fhi.pdf}
\caption{Distribution of the $M$(\hi)/\mhi$_\mathrm{lim}$\ ratio for galaxies detected in the standard, 12-h cubes (see text).}
\label{fig:detlim_fhi}
\end{figure}
The second question we wish to answer is what angular size and line-width we should assume to calculate \mhi$_\mathrm{lim}$\ for the undetected galaxies. To find an answer we study the distribution of $A$({\sc H\,i})/$A$(beam) and of the line-width $W$ of low-$\mathrm{\left(S/N\right)_{max}}$\ galaxies ($\mathrm{\left(S/N\right)_{max}}$\ $<10$). The typical low-$\mathrm{\left(S/N\right)_{max}}$\ detection is rather small on the sky. For these systems, the mode of the $A$({\sc H\,i})/$A$(beam) distribution is 1, and 2/3 of them have an area smaller than 6 beams, which we adopt as angular size for the calculation of upper limits.
We define $W$ as the line width including 80 percent of the detected emission at the sky coordinates of $\mathrm{\left(S/N\right)_{max}}$. The right panel in Fig. \ref{fig:detlim} shows that $\mathrm{\left(S/N\right)_{max}}$\ has little relation to $W$. The top $H$ symbols represent the \it FWHM \rm of the Hannning filters used to smooth the {\sc H\,i}\ cubes in the spectral direction. They demonstrate that we have properly sampled the $W$ range covered by the data. We find that median and mode of the overall $W$ distribution are $\sim 50$ km/s. Three quarters of all low-$\mathrm{\left(S/N\right)_{max}}$\ galaxies have $W$ below this value, which we adopt as line-width for the calculation of upper limits.
Based on these results, we define \mhi$_\mathrm{lim}$\ as a 3$\sigma$ signal obtained integrating over 6 beams and 50 km/s. For the typical noise of 0.5 mJy/beam at a 16 km/s velocity resolution, this corresponds to \mhi$_\mathrm{lim}$$=2.4 \cdot (d/\mathrm{10\ Mpc})^2 \times 10^6 $ {M$_\odot$}. To verify that this is a reasonable upper limit we show in Fig. \ref{fig:detlim_fhi} the histogram of the $M$(\hi)/\mhi$_\mathrm{lim}$\ ratio for the detected galaxies. For consistency, we make use of the $M$(\hi)\ value measured from the standard, 12 h-integration cube for all galaxies (this is lower than the $M$(\hi)\ value reported in Table \ref{tab:sample} for galaxies were additional {\sc H\,i}\ emission is detected in $R01$ or deep cubes). Only 6 of the 49 galaxies used for this analysis have $M$(\hi)\ below \mhi$_\mathrm{lim}$\ (NGC~2824, NGC~3182, NGC~3608, NGC~4710, NGC~5422, NGC~5866). This confirms the validity of our definition of upper limits.
As mentioned in Sec. \ref{sec:data}, we use ALFALFA spectra to determine \mhi$_\mathrm{lim}$\ for about half of all Virgo ETGs. Upper limits from ALFALFA data are calculated on one resolution element. The ALFALFA beam has a FWHM of $\sim3.5$ arcmin. Therefore, its area is larger than the 6 WSRT beams over which we calculate upper limits for the WSRT data. As above, we assume a line-width of 50 km/s. ALFALFA spectra have a typical noise of $\sim3$ mJy/beam at 11 km/s resolution, so that \mhi$_\mathrm{lim}$$=5.0 \cdot (d/\mathrm{10\ Mpc})^2 \times 10^6 $ {M$_\odot$}. The noise value of individual spectra was kindly provided by Riccardo Giovanelli.
\begin{figure}
\includegraphics[width=8.5cm]{figures/schechter3.pdf}
\caption{Normalised histogram of $M$(\hi)\ for ETGs in the ATLAS$^{\rm 3D}$\ parent sample. The dashed red line refers to {\sc H\,i}-detected ETGs. The solid red line is the corrected histogram (see text). Error bars are obtained assuming a binomial distribution. The shaded histogram represents the distribution of upper limits for undetected galaxies. The solid black line represents the best-fitting Schechter function. The dotted black line represents a Schechter function which is equal to the best-fitting one except for the slope $\alpha$, fixed to the value obtained by \protect \cite{2005MNRAS.359L..30Z}.}
\label{fig:schechter}
\end{figure}
\section{{\sc H\,i}\ mass function}
\label{sec:mf}
The large size of the \atlas\ \hi\ sample and the good sensitivity of our WSRT observations result in a relatively large number of {\sc H\,i}\ detections, making it possible to study the {\sc H\,i}\ mass function of ETGs. Figure \ref{fig:schechter} shows the distribution of $M$(\hi)\ for ETGs in the \atlas\ \hi\ sample. The red dashed line shows the observed distribution of {\sc H\,i}-detected ETGs. The shaded histogram shows the distribution of upper limits for undetected galaxies. Both distributions are normalised by the total number of galaxies in the sample.
We correct the histogram of detections for the incompleteness of our observations at low $M$(\hi). We do this by normalising the number of galaxies detected in each bin by the number of galaxies $detectable$ in that bin (i.e., the number of galaxies with \mhi$_\mathrm{lim}$\ below the upper end of the bin). This is similar to the $V/V_\mathrm{max}$ correction typically done in large flux-limited surveys \citep{1968ApJ...151..393S}. We show the corrected histogram with a red solid line in the figure. Error bars in the figure assume a binomial distribution.
The correction can be applied only to a sample of galaxies each with $M$(\hi)$\geq$\mhi$_\mathrm{lim}$. Three {\sc H\,i}-detected ETGs do not meet this criterion (NGC~3182, NGC~3499 and NGC~7332). The corrected histogram in Fig. \ref{fig:schechter} are drawn ignoring these galaxies (i.e., considering them effectively as undetected).
The corrected $M$(\hi)\ distribution of ETGs is flat between a few times $10^8$ and $\sim3\times10^9$ {M$_\odot$}, and declines rapidly above the latter value. At lower masses it exhibits a slight decline all the way to $M$(\hi)$\sim10^7$ {M$_\odot$}, where it may start rising again (but uncertainties in this mass range are large). We quantify this behaviour by fitting a Schechter function to the corrected $M$(\hi)\ distribution \citep{1976ApJ...203..297S}:
\begin{equation}
dn \propto \left(M/M^*\right)^{1+\alpha} e^{-M/M^*} d \log_{10}M,
\end{equation}
\noindent where $dn$ is the number of galaxies within a logarithmic bin of width $d\log_{10}M$. The solid black line in the figure represents the best fit, which has parameters $M^*= \left( 1.8\pm 0.7 \right) \times 10^9$ {M$_\odot$}\ and $\alpha=-0.68\pm0.16$.
The above mass function is derived neglecting {\sc H\,i}\ non-detections. Figure \ref{fig:schechter} shows that this may lead us to underestimate the mass function (and overestimate $\alpha$) below $M$(\hi)\ $\sim5\times10^7$ {M$_\odot$}. We therefore repeat the fit using only the distribution above this mass. The result is $\alpha=-0.79\pm0.23$, consistent with the fit on the full $M$(\hi)\ range. We note that the shape of the ETG {\sc H\,i}\ mass function is not driven by environmental effects. The same shape is recovered when excluding the (mostly undetected) galaxies living inside Virgo.
The value of $M^*$ for ETGs is $\sim5$ times lower than the value typically found in large, blind {\sc H\,i}\ surveys \citep[e.g.,][]{2003AJ....125.2842Z,2005ApJ...621..215S,2005MNRAS.359L..30Z,2010ApJ...723.1359M}. Both our and these previous surveys are complete at such high $M$(\hi)\ levels, so that comparing the $M^*$ values is correct. This confirms that ETGs contain smaller amounts of {\sc H\,i}\ than other galaxies.
A comparison between the low-mass end slope $\alpha$ of the ETG {\sc H\,i}\ mass function and that found by previous authors for galaxies of different morphology is complicated because of the different way samples are selected. Our sample is magnitude-limited while previous studies analyse samples of {\sc H\,i}-detected galaxies with no selection on galaxy luminosity. This means that previous studies include fainter galaxies than we do. Such galaxies contribute to the low-mass end of the {\sc H\,i}\ mass function, causing $\alpha$ to decrease. Indeed, they find $\alpha$ values of $-1.2$ to $-1.4$, i.e., the mass function keeps rising with decreasing $M$(\hi)\ below $M^*$, unlike in our ETG sample (see dotted line in Fig. \ref{fig:schechter}). We present a comparison between ETGs and spiral galaxies with the same luminosity selection in the next section. \rm
\section{Comparison to spiral galaxies}
\label{sec:spir}
In Fig. \ref{fig:KMmhi0} we compare the $M$(\hi)\ and $M$(\hi)/\lk\ distribution of ETGs to those of spiral galaxies. Red histograms refer to ETGs in the \atlas\ \hi\ sample -- solid line, dashed line and shaded histogram have the same meaning as in Fig. \ref{fig:schechter}. The blue solid line represents the distribution of spirals belonging to the ATLAS$^{\rm 3D}$\ parent sample, with the additional selection criterion $\delta \geq 10$ deg as for the ETGs studied here (Sec. \ref{sec:sample}). Spirals in the ATLAS$^{\rm 3D}$\ parent sample have the same {$M_\mathrm{K}$}\ selection as ATLAS$^{\rm 3D}$\ ETGs.
For spiral galaxies the value of $M$(\hi)\ is derived from the {\sc H\,i}\ flux available in HyperLeda \citep{2003A&A...412...57P}, assuming the distance given in Paper I. The {\sc H\,i}\ flux is available for $390/418=93$ percent of all spirals. From HyperLeda, it is not possible to know whether the remaining 28 spirals are undetected or unobserved in {\sc H\,i}. However, about 80 percent of them are fainter than {$M_\mathrm{K}$}$=-22.5$, so that they are likely to contribute to the low-$M$(\hi)\ end of the spiral distribution. Furthermore, $\sim2/3$ of them have distance larger than $\sim30$ Mpc, which makes them difficult to detect depending on the depth of the data and may explain why they are missing from the HyperLeda database. This small fraction of missing galaxies and the relatively high uncertainty on HyperLeda {\sc H\,i}\ fluxes are not particularly important for the type of comparison presented in this section.
\begin{figure*}
\includegraphics[width=18cm]{figures/histobest2.pdf}
\caption{Normalised histogram of $M$(\hi)\ (left) and $M$(\hi)/\lk\ (right) for ETGs and spirals in the ATLAS$^{\rm 3D}$\ parent sample. The dahsed red line refers to {\sc H\,i}-detected ETGs. The solid red line is the corrected histogram (see Sec. \ref{sec:mf}). The shaded histogram represents the distribution of upper limits for undetected galaxies. The solid blue line refers to spirals with available {\sc H\,i}\ flux in HyperLeda. Error bars are obtained assuming a binomial distribution.}
\label{fig:KMmhi0}
\end{figure*}
As discussed in Sec. \ref{sec:mf}, the $M$(\hi)\ distribution of ETGs is very broad between (at least) a few times $10^7$ and a few times $10^9$ {M$_\odot$}, while at lower {\sc H\,i}\ masses it is uncertain because of many upper limits. Spirals in the ATLAS$^{\rm 3D}$\ parent sample look very different, with a much narrower distribution peaking at $M$(\hi)$\sim2\times10^9$ {M$_\odot$}\ and a tail towards {\sc H\,i}\ masses below a few times $10^8$ {M$_\odot$}. The number of galaxies in this tail might be slightly higher if most spiral galaxies without available {\sc H\,i}\ flux have small $M$(\hi)\ (as argued above). However, this will not be a very large effect, and Fig. \ref{fig:KMmhi0} shows that the relative number of galaxies with $M$(\hi)\ below a few times 10$^8$ {M$_\odot$}\ is larger in ETGs than in spirals when the same {$M_\mathrm{K}$}\ selection is applied. Therefore, for {$M_\mathrm{K}$}$<-21.5$ the slope $\alpha$ of the spiral {\sc H\,i}\ mass function might be much smaller than the value of $\sim-0.7$ derived for ETGs in the previous section. Overall, 86 percent of all ETGs and 87 percent of all spirals have $M$(\hi)\ below and above $5\times10^8$ {M$_\odot$}, respectively. \rm
The $M$(\hi)/\lk\ ratio follows a peaked distribution in spirals owing to the known correlation between $M$(\hi)\ and galaxy luminosity \citep{1978A&A....68..321S}. Such correlation does not hold for ETGs, so that their $M$(\hi)/\lk\ distribution is much broader. In fact, it is approximately flat between $M$(\hi)/\lk$=10^{-3}$ and $10^{-1}$ {M$_\odot/$L$_\odot$}\ and drops to zero above this value, where the distribution of spirals peaks. We find that 85 percent of all ETGs and 92 percent of all spirals have $M$(\hi)/\lk\ below and above $10^{-2}$ {M$_\odot/$L$_\odot$}, respectively.
Early-type galaxies and spirals are distributed very differently in Fig. \ref{fig:KMmhi0}, demonstrating that ETGs, as a family, contain much less neutral hydrogen gas than spirals (a similar indication comes from the low $M^*$ value of the best-fitting Schechter function described in the previous section). However, their distributions overlap significantly in the range $M$(\hi)=$5\times10^7$ - $5\times10^{9}$ {M$_\odot$}\ and $M$(\hi)/\lk=$3\times10^{-3}$ - $10^{-1}$ {M$_\odot/$L$_\odot$}. \it A significant fraction of all ETGs contain as much {\sc H\,i}\ as spiral galaxies\rm.
This is in agreement with results of \cite{2010MNRAS.403..683C} as part of the GASS survey of massive galaxies. They show a weak relation between $M$(\hi)/$M_\mathrm{star}$ and the concentration index measured from galaxies' optical images (their Fig. 8). This index is proxy for the bulge-to-disc ratio of a galaxy \citep[e.g.,][]{2009MNRAS.394.1213W}, so that it is often used to select samples of early- or late-type systems. Adopting standard criteria for this selection, \cite{2010MNRAS.403..683C} results show that spiral galaxies are on average {\sc H\,i}-richer than ETGs, and that most {\sc H\,i}\ non-detections are ETGs for \mhi$_\mathrm{lim}$/$M_\mathrm{star}$ of a few percent. However, consistent with our result, they also show that there is substantial overlap between the gas content of early- and late-type galaxies.
Another element of similarity between spirals and ETGs with {\sc H\,i}\ is that the majority of all {\sc H\,i}-rich ETGs show disc-like {\sc H\,i}\ morphology (64 percent over the entire sample and 80 percent above $M$(\hi)=$5\times10^8$ {M$_\odot$}). Therefore, a significant population of ETGs exists which has similar $M$(\hi), $M$(\hi)/\lk\ and {\sc H\,i}\ morphology as spiral galaxies. What is then the difference between the {\sc H\,i}\ properties of these two types of galaxies?
\begin{figure}
\includegraphics[width=8.5cm]{figures/cdf_1sig2.pdf}
\caption{Distribution of {\sc H\,i}\ column density $N$(\hi). The solid red line shows the median distribution for detected ETGs. The solid blue line shows the median distribution for spirals in the WHISP sample (see text). The shaded area corresponds to the 16th and 84th percentile (equivalent to $\pm1\sigma$ for a Gaussian distribution). Thin black lines represent the best-fitting Schechter function for ETGs and spirals, with a dotted line showing the extrapolation of the best fit to column density below the sensitivity of the observations. The dashed red line represents the column density sensitivity of the shallowest total {\sc H\,i}\ images presented in Fig. \ref{fig:gallery}.}
\label{fig:nhi}
\end{figure}
To investigate this we compare the distribution of {\sc H\,i}\ column density $N$(\hi)\ in detected ETGs and spirals. For the latter we use total-{\sc H\,i}\ images constructed from WSRT datacubes as part of the WHISP survey \citep{2002ASPC..276...84V}. We make this comparison as fair as possible by: \it(i)\rm\ using WHISP images at a 30 arcsec angular resolution, similar to that of our ETG {\sc H\,i}\ images, and \it(ii)\rm\ studying only WHISP galaxies which either belong to the ATLAS$^{\rm 3D}$\ parent sample (54 galaxies), or fall within the same recessional velocity and $M_\mathrm{B}$ range as spirals in the ATLAS$^{\rm 3D}$\ parent sample but are outside the sky area covered by it (29 galaxies).
We build the normalised $N$(\hi)\ histogram for each detected ETG and spiral separately. To do so we use for each galaxy all pixels with detected {\sc H\,i}\ emission in the total {\sc H\,i}\ image. Figure \ref{fig:nhi} shows the median of all ETG histograms (red line) and of all spiral histograms (blue line). Both distributions increase with decreasing $N$(\hi)\ down to the sensitivity limit of the observations. WHISP data cover a wide range in sensitivity and are on average a factor of $\sim5$ shallower than our observations of ETGs. The relative normalisation of ETG and spiral $N$(\hi)\ distributions is therefore uncertain.
We choose to normalise the histograms in the figure to the integral of the respective best-fitting Schechter function above $N$(\hi)$=10^{19}$ cm$^{-2}$. These best fits are shown by solid black lines in the $N$(\hi)\ range where the fit was performed, and their extrapolation to lower column density is represented by a dotted black line. For spiral galaxies we find $\alpha=-0.99\pm0.04$ and $N^*=\left( 1.03\pm0.07\right)\times10^{21}$ cm$^{-2}$. For ETGs $N^*$ is an order of magnitude lower, $\left( 9.2\pm0.6\right) \times10^{19}$ cm$^{-2}$. This value is close to our sensitivity threshold (red dashed line in the figure) so that $\alpha$ is not well constrained for ETGs. Therefore, we assume the same flat slope found for spiral galaxies ($\alpha\sim-1$)\footnote{The best-fitting Schechter function to the $N$(\hi)\ distribution of spirals has a flat slope at low column densities not fully probed by the WHISP data. We verify the prediction of this fit at low column density by analysing a very deep WSRT {\sc H\,i}\ image of a prototypical nearly face-on spiral galaxy, NGC~6946 \citep{2008A&A...490..555B}. We find that the $N$(\hi)\ distribution in this galaxy is flat and in excellent agreement with our best-fitting spiral Schechter function between a few times $10^{18}$ and $\sim10^{20}$ cm$^{-2}$, confirming the validity of our fit.}. We note that, although in different contexts, several authors have successfully parameterised the distribution of $N$(\hi)\ in galaxies using a Schechter function \citep[e.g.,][]{2005MNRAS.364.1467Z}.
Figure \ref{fig:nhi} and the above Schechter function parameters show a significant difference between ETGs and spirals. The $N$(\hi)\ distribution of spirals is very broad and stretches up to a few times $10^{21}$ cm$^{-2}$. In contrast, the distribution of $N$(\hi)\ in ETGs drops very quickly and, on average, stops at $\sim3\times10^{20}$ cm$^{-2}$. Most ETGs never reach column densities above $5\times10^{20}$ cm$^{-2}$ at the resolution of our data. This value is the mean column density of {\sc H\,i}\ within the bright disc of spiral galaxies, derived from the tight relation between {\sc H\,i}\ mass and {\sc H\,i}\ radius within the 1 {M$_\odot$}/pc$^2=1.25\times 10^{20}$ cm$^{-2}$ isophote \citep[e.g.,][]{1997A&A...324..877B,2005A&A...442..137N}. Therefore, ETGs rarely host neutral atomic gas as dense as even the \it average \rm column density in the disc of spirals. It is interesting to note that the few ETGs with {\sc H\,i}\ at about 10$^{21}$ cm$^{-2}$ -- NGC~2685, NGC~2764, NGC~3619 and NGC~7465 -- are all galaxies with large amounts of molecular gas (Paper IV), complex and prominent dust distributions and star-forming regions (Paper II).
So we find that, as far as {\sc H\,i}\ properties are concerned, the main difference between ETGs with large amounts of {\sc H\,i}\ and spirals is that the former miss the high-column-density {\sc H\,i}\ typical of the bright stellar disc of the latter. This is reasonable as the star-formation rate per unit area is much larger in spirals than in ETGs. Instead, the {\sc H\,i}\ found in ETGs has column densities similar to those observed in the outer regions of spirals. In spirals, gas in these regions often exhibits warps and unsettled morphology and kinematics, and is taken as a signature of the on-going accretion of gas on the host galaxy \citep{2008A&ARv..15..189S}. This is similar to what we find in ETGs. As mentioned in Sec. \ref{sec:morph}, most of the {\sc H\,i}-rich ETGs have a disc- or ring-like morphology, but in many cases part of the gas has not settled yet (or has been disturbed recently). This suggests that despite appearing as extremely different objects in their central regions, {\sc H\,i}-rich ETGs and spirals may look very similar in their outskirts.
Based on numerical simulations, several authors have argued that it is possible to form an ETG with a disc by merging two gas-rich galaxies \citep[e.g.,][]{2001ApJ...555L..91N,2002MNRAS.333..481B,2005ApJ...622L...9S,2006MNRAS.372..839N,2006ApJ...641...21R}. The distribution of the gas around the simulated merger remnants resembles in many aspects the {\sc H\,i}\ systems discussed here. The size and mass of the distribution varies greatly depending on the properties of the merging galaxies and the merger geometry, and evolves with time since the merging. Both gas discs and rings can form as a result of these events, and gaseous tidal tails can extend to large distance from the remnant and be re-accreted at a later time. Our results suggest that, in order for the remnant to be an ETG, gas must settle on these discs in such a way to keep its column density low. Future comparison to simulations could explore this aspect in more details.
\section{Relation between {\sc H\,i}, galaxy luminosity and environment}
\label{sec:env}
\begin{figure*}
\includegraphics[width=18cm]{figures/sixpan.pdf}
\caption{$M$(\hi)\ (top) and $M$(\hi)/\lk\ (bottom) plotted against galaxy {$M_\mathrm{K}$}\ (left), $\Sigma_{10}$ (middle), and $\Sigma_3$ (right). Different symbols represent different {\sc H\,i}\ morphologies as explained in the legend. Red and blue markers represent galaxies inside and outside the Virgo cluster, respectively.}
\label{fig:mhi}
\end{figure*}
In previous sections we describe the distribution of ETG {\sc H\,i}\ morphology, $M$(\hi)\ and $M$(\hi)/\lk. In this section we investigate whether these properties depend on galaxy luminosity and environment.
Figure \ref{fig:mhi} shows $M$(\hi)\ (top) and $M$(\hi)/\lk\ (bottom) of all galaxies in the \atlas\ \hi\ sample as a function of galaxy magnitude $M_\mathrm{K}$ (left), large-scale environment density $\Sigma_{10}$ (middle) and local environment density $\Sigma_3$ (right). Values of $M_\mathrm{K}$ are listed in Paper I. Values of $\Sigma_{10}$ and $\Sigma_3$ are listed in Paper VII (see Sec. \ref{sec:sample} for the definition of these parameters).
In the figure, triangles indicate upper limits on $M$(\hi). Other markers represent different {\sc H\,i}\ morphologies as explained in the legend on the top-right panel (see Sec. \ref{sec:morph} for a definition of {\sc H\,i}\ morphological classes). Red and blue markers represent galaxies inside and outside the Virgo cluster, respectively.
We note that the \atlas\ \hi\ sample shows no relation between {$M_\mathrm{K}$}\ and $\Sigma_{10}$ or $\Sigma_{3}$. Therefore, trends with environment can be separated from trends with galaxy luminosity.
\subsection{{\sc H\,i}\ and galaxy luminosity}
There is no strong trend of {\sc H\,i}\ properties with galaxy luminosity. Figure \ref{fig:mhi} hints to a lack of {\sc H\,i}\ at bright {$M_\mathrm{K}$}. This is in agreement with results by \cite{2010MNRAS.403..683C}, although a comparison of our results is not straightforward because of the different sample selections. In our case however, the small number of bright galaxies makes this trend relatively weak from a statistical point of view. For example, we consider the two sub-samples of galaxies above and below $M$(\hi)=$5\times10^8$ {M$_\odot$}, respectively. According to a two-sample KS test, the hypothesis that their {$M_\mathrm{K}$}\ distributions are drawn from a same parent distribution is rejected just at the $\sim 93$ percent confidence level.
Massive galaxies appear to exhibit a disturbed {\sc H\,i}\ morphology more frequently than fainter systems. The fraction of $u$'s is $7/33=21\pm8$ percent for the brightest 20 percent of the \atlas\ \hi\ sample, and $7/133=5\pm2$ percent for the remaining galaxies. This seems to indicate that recent (major or minor) interaction with other gas-rich galaxies is relatively common for massive ETGs. For example, Paper IX discuss evidence of the major-merger origin of 2 of these massive $u$ galaxies (NGC~0680 and NGC~5557) based on the {\sc H\,i}\ data presented here and on deep optical imaging, and argue that the formation event must have occurred after $z\sim0.5$. However, while {\sc H\,i}\ observations can reveal these past interactions, in a large fraction of these systems the {\sc H\,i}\ is distributed at large distance from the galaxy so that it is not clear whether in all cases gas is or will be accreted on the central region of the ETG (e.g., NGC~3193 and NGC~5557 in Fig. \ref{fig:sequenceb}).
Figure 8 suggests also that the fraction of $d$'s is higher in fainter galaxies: $4/33=12\pm6$ percent for the faintest 20 percent of the \atlas\ \hi\ sample, against $6/133=5\pm2$ percent for the rest of the galaxies. However, within the error bars this difference is only marginal.
\subsection{{\sc H\,i}\ mass and environment}
\label{sec:envelope}
Several previous authors reported a strong difference in {\sc H\,i}\ detection rate between ETGs living inside and outside the Virgo cluster. Within the \atlas\ \hi\ sample we detect $4/39=10\pm5$ percent of all ETGs in Virgo and $49/127=39\pm6$ percent of all ETGs outside Virgo. This is in good agreement with results from \citeauthor{2007A&A...474..851D} (\citeyear{2007A&A...474..851D}; $2/32=6\pm4$ percent inside Virgo for $M_\mathrm{B}\leq-18$, comparable to our {$M_\mathrm{K}$}\ selection), \citeauthor{2009A&A...498..407G} (\citeyear{2009A&A...498..407G}; $5/14=36\pm16$ percent outside Virgo with the same $M_\mathrm{B}$ selection) and O10 on the Sauron sample ($1/13=8\pm8$ percent inside and $14/20=70\pm19$ percent outside Virgo; the latter is slightly higher than our result but the typical distance of galaxies in the Sauron sample is lower and, unilke O10, our detection rate does not include a number of cases where it is unclear whether {\sc H\,i}\ belongs to the observed ETG or not -- see Fig. \ref{fig:gallery_undet_a}). Our result represents a significant improvement on the statistical significance of these detection rates.
While we confirm the Virgo vs. non-Virgo dichotomy, we note that environmental effects are more subtle than this. Firstly, $M$(\hi)\ and $M$(\hi)/\lk\ vary smoothly with environment density rather than showing clear breaks and sharp transitions. Figure \ref{fig:mhi} suggests the presence of an envelope of decreasing $M$(\hi)\ and $M$(\hi)/\lk\ with increasing environment density over the entire density range covered by our sample.
One cause of this envelope is that the {\sc H\,i}-richest ETGs live in the poorest environment. To investigate the significance of this result we compare the $\Sigma_{10}$ distributions of galaxies with $M$(\hi)/\lk\ above and below $3\times10^{-2}$ {M$_\odot/$L$_\odot$}. A two-sample KS test rejects the hypothesis that the two distributions are the same at the 99.98 percent confidence level (99.74 percent if we exclude galaxies in Virgo from this analysis). Similarly, the $\Sigma_{3}$ distributions of the same two samples are different at the 99.6 percent confidence level.
We find a similar result in our study of the CO content of ETGs (Paper IV) -- the most CO-rich ETGs live in the poorest environments. Therefore, the observation of both {\sc H\,i}\ and H$_2$ suggests that ETGs living in low-density environment accrete more cold gas (possibly from the surrounding medium) and/or can retain it for a longer period of time. This can be caused by the lack of both large neighbouring galaxies and a hot medium in such poor environments.
Another clue to the different gas-accretion histories of ETGs living in different environments might be the way star formation activity changes as a function of environment density. We divide the full ATLAS$^{\rm 3D}$\ sample in four $\Sigma_{10}$ bins with bin edges at $\mathrm{log_{10}} \ \Sigma_{10}/$Mpc$^2=-2,-1,0,1,2$. Moving from the poorest to the richest environment we find dust/blue features (indicative of star formation; see Sec. \ref{sec:h2}), in $21/65$, $18/105$, $6/47$ and $6/43$ of all galaxies. We conclude that star formation signatures are more frequent at the poorest environment ($32\pm7$ percent) than in all the other $\Sigma_{10}$ bins ($\sim15\pm5$ percent in each of them separately).
Variations in the physical properties of the medium (e.g., density and temperature) and in the rate of galaxy interaction might be causing the gradual decrease of $M$(\hi)\ and $M$(\hi)/\lk\ with environment density. Such transition in {\sc H\,i}\ properties is visible also within the Virgo cluster, suggesting that galaxies living at the outskirts of Virgo are different from those closer to its centre.
\subsection{Centre and outskirts of the Virgo cluster}
Galaxies detected in Virgo are NGC~4262, NGC~4406, NGC~4694 and NGC~4710. Their projected distance from M87 is 1.0, 0.5, 1.3 and 1.6 Mpc, respectively. Surface-brightness fluctuation distances are available for NGC~4262 and NGC~4406, and place them at a distance of 1.8 and 0.4 Mpc from M87 along the line of sight, respectively (Paper I). NGC~4406 is the only galaxy detected in {\sc H\,i}\ close to the cluster centre, so that the {\sc H\,i}\ detection rate in this region is just $1/26=4\pm4$ percent. In contrast, the detection rate outside 1 Mpc from M87 is $3/13=23\pm13$ percent. So we find that {\sc H\,i}\ detection rate, $M$(\hi)\ and $M$(\hi)/\lk\ (see Fig. \ref{fig:mhi}) in the centre of Virgo are significantly lower than outside the cluster, while Virgo's outskirts appear as a transition region.
In Paper VII we find that the variation of galaxy morphological mix with environment density also continues within Virgo. In particular, the morphology-density relation changes slope at the Virgo cluster centre, where all Virgo slow rotators are found. This is in agreement with the above {\sc H\,i}\ result and strengthens the conclusion that different processes for the formation of ETGs must be at work inside the cluster core. These processes have the effect of generating a population of ETGs which is poorer of {\sc H\,i}\ and significantly richer of slow rotators than galaxies in less dense environments, including the cluster outskirts.
The above results are in good agreement with conclusions from the VIVA {\sc H\,i}\ survey of spirals in Virgo by \cite{2009AJ....138.1741C}. They find very {\sc H\,i}-rich spirals with a gas disc larger than the optical disc (typical of field spirals) only further than 1 Mpc from M87 in projection. Closer to the cluster centre they find only galaxies whose {\sc H\,i}\ disc is truncated to the size of the stellar disc (or even smaller for systems very close to M87), and galaxies with gas tails caused by a combination of ram-pressure stripping and tidal interaction with neighbouring galaxies. The 1 Mpc cluster-centric radius around which \cite{2009AJ....138.1741C} see a transition in the {\sc H\,i}\ properties of spirals matches our result on the variation of ETG {\sc H\,i}\ properties within Virgo.
\begin{figure*}
\includegraphics[width=34mm]{figures/HISPECIALfigures/UGC06176.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/UGC09519.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC6798.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC2594.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC4150.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC7332.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC5582.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC7280.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/UGC03960.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC3182.pdf}
\caption{The ten {\sc H\,i}\ detections with the lowest $\Sigma_{3}$. Images are sorted according to increasing $\Sigma_3$, left to right, top to bottom. We refer to the caption of Fig. \ref{fig:gallery} for a description of the content of each image. In this figure the top-right scale bar indicates 10 kpc at the galaxy distance}
\label{fig:poorenv}
\end{figure*}
\begin{figure*}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC4026.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC5198.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC4111.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC3998.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC4521.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC4694.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC3384.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC4262.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC0680.pdf}
\includegraphics[width=34mm]{figures/HISPECIALfigures/NGC4406.pdf}
\caption{The ten {\sc H\,i}\ detections with the highest $\Sigma_{3}$. Images are sorted according to increasing $\Sigma_3$, left to right, top to bottom. We refer to the caption of Fig. \ref{fig:gallery} for a description of the content of each image. In this figure the top-right scale bar indicates 10 kpc at the galaxy distance}
\label{fig:richenv}
\end{figure*}
Neutral hydrogen is known to be quickly removed from galaxies as they move within the hot intra-cluster medium \citep[e.g.,][]{1972ApJ...176....1G}. This is indeed the common interpretation of the fact that Virgo ETGs, which are a virialised population and have therefore completed many orbits within the cluster, are so {\sc H\,i}-poor \citep[e.g.,][O10]{2007A&A...474..851D}. Because of these considerations, the relatively high $M$(\hi)\ and {\sc H\,i}\ detection rate of ETGs in the outskirts of Virgo suggests that at least a fraction of them is falling now for the first time in the cluster (a similar argument is commonly used to claim that Virgo spirals are making their first passage through the cluster). This implies that at least a fraction of all Virgo galaxies may have joined the cluster already with an early-type morphology, rather than having become ETGs within the cluster. Their current morphology may result from pre-processing inside galaxy groups, as suggested also by our result that the main driver of galaxies' morphology-density relation are processes occurring at the galaxy-group scale (Paper VII).
The {\sc H\,i}\ properties of ETGs in Virgo are in contrast with results of our CO observations (Paper IV). Unlike {\sc H\,i}-rich objects, ETGs with molecular gas do not avoid the cluster centre. Their detection rate is $5/33=15\pm7$ percent within 1 Mpc from M87 and $6/25=24\pm10$ percent farther than 1 Mpc from M87 (for comparison, it is $47/202=23\pm3$ percent outside Virgo). Given the errorbars, it is still possible that the CO detection rate is lower in the centre of Virgo than at its outskirts or outside the cluster. However, it is clear that the difference between ETGs living in different environments is much more obvious in {\sc H\,i}\ than in H$_2$ (see \citealt{1989ApJ...344..171K} for a similar result on spiral galaxies).
The H$_2$ mass of ETGs inside the cluster is also consistent with that of galaxies outside it (with the aforementioned exception of the very H$_2$-rich galaxies in the poorest environment; see Sec. \ref{sec:envelope}). This implies that the lack of {\sc H\,i}\ in Virgo is related to a high $M$(H$_2$)/$M$(\hi)\ ratio rather than an overall gas deficiency. A possible explanation is that {\sc H\,i}\ is more easily stripped than H$_2$ when a galaxy moves through a hot medium. This is in general true over the full extent of a galaxy because {\sc H\,i}\ is typically distributed out to larger radii than CO and is therefore less bound to the galaxy \citep{2009ApJ...697.1811F}. This situation is particularly frequent in ETGs, where the {\sc H\,i}\ often extends to tens of kpc from the galaxy (see Fig. \ref{fig:gallery}). However, many ETGs outside the cluster contain {\sc H\,i}\ in their central regions, so the question is why we do not detect such central {\sc H\,i}\ (but do detect central CO) in Virgo.
Neutral hydrogen in the centre of ETGs might be more easily stripped than CO because of its lower surface density \citep{2011arXiv1106.2554Y}. Another important indication is that ETGs with CO appear as a virialised population within Virgo (Paper IV), so that they may have completed many orbits within the cluster. A prolonged residence within the cluster might be the reason why these galaxies are so {\sc H\,i}\ poor. They are not only stripped of their gas, but the lack of other {\sc H\,i}-rich companions means that no further {\sc H\,i}\ accretion can occur (see O10). While this seems a plausible picture, we note that at least in some cases CO is stripped together with the {\sc H\,i}\ from the centre of galaxies \citep{2008A&A...491..455V}, so that we do not have a definitive answer to this problem yet.
\subsection{{\sc H\,i}\ morphology-density relation}
In Sec. \ref{sec:envelope} we show that galaxies in poor environment can reach higher {\sc H\,i}\ masses than those found at higher density. These objects also tend to have a more regular {\sc H\,i}\ morphology than gas-rich ETGs in denser environments. We illustrate the relation between {\sc H\,i}\ morphological mix and environment density in Figs. \ref{fig:poorenv} and \ref{fig:richenv}. These show total {\sc H\,i}\ images of the ten detections with the lowest and highest $\Sigma_3$, respectively, sorted according to increasing $\Sigma_3$ (left to right, top to bottom). Figure \ref{fig:richenv} includes 3/4 of all Virgo detections, all 3 Ursa Major detections (NGC~3998, NGC~4026 and NGC~4111; {\sc H\,i}\ in these galaxies is discussed also by \citealt{2001ASPC..240..867V}) and 4 galaxies living in relatively rich groups. The fraction of settled {\sc H\,i}\ systems (i.e., $D$ and $d$) is high at the lowest environment densities: $8/10=80\pm28$ percent, with the only $u$ system possibly being a misclassified $d$ (see Sec. \ref{sec:morph}). In contrast, it is very low at the highest densities: $2/10=20\pm14$ percent. At high $\Sigma_3$ most {\sc H\,i}-rich systems look disturbed (including some classified as $D$, like NGC~3998).
The trend of {\sc H\,i}\ morphology with environment density is stronger using $\Sigma_3$ than $\Sigma_{10}$. For example, galaxies with a very disturbed {\sc H\,i}\ morphology figure among the ten objects with the lowest $\Sigma_{10}$ (NGC~1023, NGC~7465); and the fraction of settled {\sc H\,i}\ systems at the lowest $\Sigma_{10}$ is the same as that at the highest $\Sigma_{10}$ ($5/10$ against $4/10$). This indicates that processes influencing the {\sc H\,i}\ morphology occur on a galaxy-group scale, and that a disturbed {\sc H\,i}\ morphology is more clearly related to the presence of close neighbours than to an overall, large-scale environment over-density.
It is interesting that a similar conclusion was reached in Paper VII based on the morphology-density relation of galaxies within the ATLAS$^{\rm 3D}$\ volume. There we found that processes occurring on a galaxy-group scale must be the main driver of the relation, which is tighter and steeper using $\Sigma_3$ rather than $\Sigma_{10}$. The variation of {\sc H\,i}\ morphology as a function of local environment density confirms the importance of such processes for the evolution of ETGs.
\section{Summary and Conclusions}
\label{sec:summ}
We study the {\sc H\,i}\ properties of all 166 ETGs in the volume-limited ATLAS$^{\rm 3D}$\ sample above $\delta=10$ deg and further than 15 arcmin from Virgo A. The sample includes galaxies within 42 Mpc and brighter than {$M_\mathrm{K}$}=$-21.5$. This is the largest sample of ETGs with deep, interferometric {\sc H\,i}\ observations to date.
We confirm earlier findings that the {\sc H\,i}\ detection rate of ETGs depends strongly on environment density \citep[][O10]{2007A&A...474..851D,2009A&A...498..407G}. We detect $4/39=10\pm5$ percent of all ETGs inside the Virgo galaxy cluster, and $49/127=39\pm6$ percent of all ETGs outside it. This is consistent with results from previous authors, although on a stronger statistical basis. The lower detection rate of ETGs in {\sc H\,i}\ surveys of larger volumes is easily explained by their much poorer sensitivity. For example the $M$(\hi)\ detection limit in the study of \cite{2011ApJ...732...92T} is above $10^9$ {M$_\odot$}\ in most of the surveyed volume. Figure \ref{fig:KMmhi0} shows that, at such sensitivity, their survey is expected to detect only $\sim10$ percent of all ETGs.
We classify galaxies according to their {\sc H\,i}\ morphology. We find that 1/5 of all ETGs (and 1/4 of all ETGs outside Virgo) host {\sc H\,i}\ distributed on a regularly rotating disc or ring. These systems represent the majority of the {\sc H\,i}\ detections, 64 percent, so that if an ETG is detected in {\sc H\,i}\ it will most likely host a gas disc or ring. Another 8 percent of all ETGs (26 percent of all detections) host {\sc H\,i}\ in an unsettled configuration, while the remaining detected galaxies are surrounded by {\sc H\,i}\ clouds scattered around the stellar body and not obviously associated to it.
Although most detections can be easily classified within this scheme, we do find a number of galaxies intermediate (or in transition) between classes. Indeed, we claim that the {\sc H\,i}\ morphology of ETGs varies in a continuous way, going from very regular discs and rings to unsettled gas distributions, and from these to systems of scattered clouds which may or may not be associated to the host galaxy. This continuity may be related to the time passed since the last major episode of gas accretion or stripping relative to the {\sc H\,i}\ orbital time in each galaxy.
We divide regular, rotating {\sc H\,i}\ systems in two classes based on their size relative to the stellar body of the host galaxy: $D$ (large discs/rings -- 24 galaxies) and $d$ (small discs -- 10 galaxies). The former contain between $10^8$ and $10^{10}$ {M$_\odot$}\ of gas distributed out to tens of kpc from the stellar body. In half of these systems the {\sc H\,i}\ is morphologically or kinematically misaligned with respect to the stellar body. In contrast, $d$'s contain typically less than $10^8$ {M$_\odot$}\ of {\sc H\,i}\ confined within the stellar body and morphologically and/or kinematically aligned to it in nearly all cases.
We investigate the role of {\sc H\,i}\ in fuelling star formation within the host ETG. Confirming results by O10, we find that the detection rate of signatures of star formation (molecular gas, dust discs/filaments and blue features) is high ($\sim70$ percent) for ETGs with {\sc H\,i}\ within the stellar body. Namely, such features are found in all $d$'s, $\sim60$ percent of al $D$'s with central {\sc H\,i}\ and $\sim50$ percent of all unsettled {\sc H\,i}\ systems with central {\sc H\,i}. On the contrary, they are found in just $\sim15$ percent of all ETGs with no central {\sc H\,i}\ (or no {\sc H\,i}\ at all at the sensitivity of our observations), highlighting the role of {\sc H\,i}\ in enriching ETGs with material for star formation.
The ISM in the centre of ETGs is dominated by molecular gas, whose mass is a factor of a few up to $\sim100$ larger than the {\sc H\,i}\ mass. Galaxies hosting a small {\sc H\,i}\ disc ($d$) reach particularly high $M$(\hi)/$M$(H$_2$)\ values (larger than 10 in half of the cases), suggesting that they are forming molecular gas at efficiencies comparable to those of spirals.
We parameterise the {\sc H\,i}\ mass function of ETGs with a Schechter function with $M^*= \left( 1.8\pm 0.7 \right) \times 10^9$ {M$_\odot$}\ and $\alpha=-0.68\pm0.16$. The value of $M^*$ is $\sim5$ times lower than typical values found for samples dominated by spiral galaxies, confirming that ETGs are, as a a family, poorer of {\sc H\,i}. The value of the faint-end slope $\alpha$ means that the mass function for a {$M_\mathrm{K}$}-limited sample decreases with decreasing $M$(\hi)\ below $M^*$.
We compare the {\sc H\,i}\ properties of ETGs to those of spirals. The $M$(\hi)\ and $M$(\hi)/\lk\ distributions of ETGs are very broad, reflecting the large variety of {\sc H\,i}\ content of these galaxies, and confirming the lack of correlation between ETG {\sc H\,i}\ mass and luminosity. The distributions of spirals are much narrower and shifted towards larger $M$(\hi)\ and $M$(\hi)/\lk. Yet, we find a substantial overlap in the $M$(\hi)\ and $M$(\hi)/\lk\ distributions of ETGs and spirals (consistent with \citealt{2010MNRAS.403..683C}). A significant fraction of all ETGs can have as much {\sc H\,i}\ as spiral galaxies, and in the majority of the cases this is distributed on a large, rotating disc or ring.
We investigate the difference between such {\sc H\,i}-rich ETGs and spiral galaxies. We find that in the former the {\sc H\,i}\ column density is typically very low. Early-type galaxies rarely reach the gas density typical of the bright stellar disc of spiral galaxies, consistent with their lower star formation rate per unit area. The few galaxies that do, show clear signs of star formation and prominent dust features. Gas column density appears therefeore as the decisive factor in determining whether a galaxy is of early or late morphological type.
We find a minor dependence of ETG {\sc H\,i}\ properties on galaxy luminosity. Very luminous galaxies seem to contain less {\sc H\,i}, but the only clear result is that {\sc H\,i}\ in these systems is typically found in an unsettled configuration. This may reflect a higher rate of interaction with gas-rich companions for massive ETGs compared to fainter objects, although it is not clear how much of this gas is eventually accreted on the centre of the host galaxy.
We do find clear trends of {\sc H\,i}\ properties with environment. These go well beyond the known Virgo-vs.-non Virgo dichotomy mentioned at the beginning of this section. We find a smooth envelope of decreasing $M$(\hi)\ and $M$(\hi)/\lk\ with environment density. Consistent with results from the ATLAS$^{\rm 3D}$\ CO survey presented in Paper IV, we find that the gas-richest galaxies live in poor environments. This indicates that the cold-gas accretion rate is higher at these densities, and/or that these galaxies can retain their gas reservoir for longer periods of time. This effect is likely driven by a combination of a relatively quiet merging history and by the lack of a hot medium, and is accompanied by a higher fraction of objects exhibiting signs of recent star-formation in this environment.
The gradual decrease of {\sc H\,i}\ mass with environment density continues within the Virgo cluster. Consistent with results from the study of spiral galaxies \citep{2009AJ....138.1741C}, we find that the {\sc H\,i}\ properties of ETGs living at the outskirts of Virgo are intermediate between those of galaxies outside Virgo and those of the very {\sc H\,i}-poor objects in its central region. Being {\sc H\,i}-rich, galaxies at the Virgo outskirts are likely falling in the cluster for the first time, and they do so already with an early-type morphology. This suggests that pre-processing on a galaxy-group scale is fundamental for the evolution of ETGs, as concluded also in Paper VII.
Galaxies inside $\sim1$ Mpc from the centre of the cluster are significantly poorer of {\sc H\,i}\ than galaxies outside Virgo, with a drop in the {\sc H\,i}\ detection rate to just a few percent. This agrees with the observation of truncated {\sc H\,i}\ discs in spiral galaxies in this region, and resonates with our finding that the fraction of slow rotating ETGs, which is very low at all environment densities, increases in the core of Virgo. This highlights that different processes must drive the evolution of galaxies deep inside the cluster, among which the presence of a hot medium (and therefore lack of {\sc H\,i}).
Unlike {\sc H\,i}, molecular gas can be found in ETGs living deep inside Virgo. These systems have $M$(H$_2$)\ similar to that of galaxies outside Virgo, so that the lack of {\sc H\,i}\ seems to imply a high molecular-to-atomic mass ratio. A possible interpretation is that molecular gas is more difficult to remove via ram-pressure stripping than {\sc H\,i}, so that ETGs that have long been in the cluster (and therefore long been stripped of their {\sc H\,i}) can still host molecular gas deep in the centre of their gravitational potential (Paper IV). However, it is not clear why {\sc H\,i}\ is not detected in the centre of these galaxies, as observations show that CO can be stripped together with {\sc H\,i}\ in at least some cases \citep{2008A&A...491..455V}.
We find that the {\sc H\,i}\ morphology also depends on environment density. In poor environment the typical {\sc H\,i}\ detection exhibits a large, settled {\sc H\,i}\ disc or ring. These systems must have been in place for many Gyr and indicate that very recent accretion/merging events have been minor. On the contrary, disturbed and unsettled {\sc H\,i}\ morphologies are very frequent at higher environment density. Such {\sc H\,i}\ morphology-density relation is clearer when measuring environment density on a local scale, indicating that processes occurring on a galaxy-group scale are driving factors for the evolution of ETGs. This agrees with the study of the kinematical morphology-density relation presented in Paper VII, where we conclude that processes occurring on this scale determine the increase of the ETG fraction (and decrease of the spiral fraction) with environment density.
As a concluding remark, we have demonstrated that a large fraction of ETGs contain significant amounts of {\sc H\,i}. It is quite normal for these galaxies to host neutral hydrogen gas, and {\sc H\,i}-rich ETGs should not be regarded as odd or rare systems. On the contrary, {\sc H\,i}\ seems to be playing an important role in replenishing ETGs with cold gas and fuelling the formation of new stars and a stellar disc in these systems. The amount and morphology of this {\sc H\,i}\ depend weakly on galaxy luminosity, one of the main parameters determining some of the most important galaxy properties such as their structure and stellar populations. Even trends with environment density, another fundamental parameter for galaxy evolution, are characterised by a very large scatter. This scatter suggests that the gas accretion history of ETGs varies widely from galaxy to galaxy even at fixed galaxy mass and environment density. A logical consequence of this large scatter is that ETGs are not such because they lack {\sc H\,i}\ -- many of them contain plenty of it. Instead we have shown that most ETGs must accrete gas in such a way that it remains at low column density -- too low to support the higher levels of star formation seen in spiral galaxies.
\section*{Acknowledgments}
PS would like to thank Mike Sipior for his help setting up the WSRT data reduction pipeline at ASTRON, Riccardo Giovanelli for providing ALFALFA spectra, and Gyula J{\'o}zsa for scheduling the WSRT observations.
MC acknowledges support from a Royal Society University Research Fellowship.
This work was supported by the rolling grants `Astrophysics at Oxford' PP/E001114/1 and ST/H002456/1 and visitors grants PPA/V/S/2002/00553, PP/E001564/1 and ST/H504862/1 from the UK Research Councils. RLD acknowledges travel and computer grants from Christ Church, Oxford and support from the Royal Society in the form of a Wolfson Merit Award 502011.K502/jd. RLD also acknowledges the support of the ESO Visitor Programme which funded a 3 month stay in 2010.
SK acknowledges support from the the Royal Society Joint Projects Grant JP0869822.
RMcD is supported by the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc., on behalf of the international Gemini partnership of Argentina, Australia, Brazil, Canada, Chile, the United Kingdom, and the United States of America.
TN and MBois acknowledge support from the DFG Cluster of Excellence `Origin and Structure of the Universe'.
MS acknowledges support from a STFC Advanced Fellowship ST/F009186/1.
NS and TD acknowledge support from an STFC studentship.
MBois has received, during this research, funding from the European Research Council under the Advanced Grant Program Num 267399-Momentum.
The authors acknowledge financial support from ESO.
\bibliographystyle{mn2e}
|
\section*{Introduction}
Since its launch in 2008, the Large Area Telescope (LAT) on board the
Fermi Gamma-ray Space Telescope \cite{aaa+09c} has detected
whole populations of objects previously unseen in the
$\gamma$-ray band. These include globular clusters (GCs), which are ancient
spherical groups of $\sim 10^5$ stars held together by their mutual
gravity. As a class, their $\gamma$-ray spectra show evidence for an
exponential cut-off at high energies
\cite{aaa+10c,tkh+11}, a characteristic signature of
magnetospheric pulsar emission. This is not surprising because radio
surveys have shown that GCs contain large numbers of pulsars
\cite{footnote1}, neutron stars that emit radio and
in some cases X-ray and $\gamma$-ray pulsations.
The first GC detected at $\gamma$-ray energies was
47~Tucanae \cite{aaa+09a}, soon followed by Terzan~5 \cite{khc10}
and nine others \cite{aaa+10c,tkh+11}. Even so, no individual
pulsars in these clusters were firmly identified in
$\gamma$-rays\cite{footnote2}.
GCs are more distant than most $\gamma$-ray pulsars observed in the
Galactic disk \cite{aaa+10}, thus most pulsars in them should be too
faint to be detected individually. The Fermi\ LAT lacks the spatial
resolution required to resolve the pulsars in GCs, which
tend to congregate within the inner arcminute of the cluster.
Hence, $\gamma$-ray photons emitted by all pulsars in a given GC
increase the photon background in the folded $\gamma$-ray profiles
of each individual pulsar in that cluster.
One of the GCs detected at $\gamma$-ray energies is
NGC~6624 \cite{tkh+11}, located at a distance $d = 8.4 \pm 0.6$\,kpc
from Earth
\cite{vfo07}. With a radio flux density at 400 MHz
of $S_{400} = 16$ mJy \cite{bbl+94}, J1823$-$3021A\, is the brightest of the six pulsars
known in the cluster. It has been regularly timed with
the Jodrell Bank and Parkes radio telescopes since discovery, and
with the Nan\c{c}ay radio telescope since the
launch of the Fermi\ satellite. The resulting radio ephemeris
(Table S1) describes the measured pulse times of arrival very well
for the whole length of the Fermi\ mission, the root mean square of
the timing residuals being 0.1\% of the pulsar rotational period.
Thus we can confidently use it to assign a pulsar
spin phase $\phi$ to every $\gamma$-ray ($>$0.1\,GeV) photon arriving at the
Fermi-LAT from the direction (within 0.8$^{\circ}$) of the
pulsar. We selected photons that occurred between 4 August 2008 and
4 October 2010 that pass the ``Pass 6 diffuse''
$\gamma$-ray selection cuts \cite{aaa+09c}.
The resulting pulsed $\gamma$-ray signal (above 0.1\,GeV, Fig.~1)
is very robust, with an H-test value of 64 \cite{db10} corresponding to
6.8\,$\sigma$ significance. The data are well modeled
by a power law with spectral index $1.4 \pm 0.3$\ and an exponential
cutoff at an energy of $1.3 \pm 0.6$\,GeV, typical of the values found
for other $\gamma$-ray pulsars [see supporting online material (SOM)].
The two peaks are aligned, within uncertainties, with the two main
radio components at spin phases $\phi_1 = 0.01 \pm 0.01$ and
$\phi_2 = 0.64 \pm 0.01$ (Fig.~1).
The pulsed flux above
0.1\,GeV, averaged over time, is $F_{\gamma}\,=\,$$(1.1 \pm 0.1 \pm 0.2) \times 10^{-11}$\,erg\,cm$^{-2}$\,s$^{-1}$,
where the first errors are statistical and the second are systematic
(SOM). The large distance of NGC~6624 implies that
J1823$-$3021A\ is one of the most distant $\gamma$-ray pulsars detected \cite{aaa+10}.
This makes it the most luminous $\gamma$-ray MSP
to date \cite{LAT_MSPs}: Its total emitted power is $L_{\gamma} = 4 \pi
d^2 f_{\Omega} F_{\gamma}\,=\,$$(8.4 \pm 1.6 \pm 1.5) \times 10^{34}$\,$(f_{\Omega}/0.9)\,$erg\,s$^{-1}$.
We obtained the statistical uncertainty by adding the uncertainties
of $d$ and $F_{\gamma}$ in quadrature. The term $f_{\Omega}$ is
the power per unit surface across the whole sky divided by power per
unit surface received at Earth's location; detailed modeling
of the $\gamma$ and radio light curves provides a best fit centered at
0.9, but with a possible range from 0.3 to 1.8 (SOM).
The LAT image of the region around NGC~6624 during the on-pulse
interval ($0.60<\phi<0.67$ and $0.90<\phi<1.07$) shows a bright
and isolated $\gamma$-ray source that is consistent with the location
of J1823$-$3021A\ (Fig.~\ref{fig:NGC6624_image});
in the off-pulse region ($0.07<\phi<0.60$ and $0.67<\phi<0.90$)
no point sources in the energy band 0.1 - 100\,GeV are
detectable.
Assuming a typical pulsar spectrum with a spectral index of 1.5 and a
cut-off energy of 3\,GeV, we derived, after scaling to the full pulse
phase, a 95\% confidence level upper limit on the point source energy
flux of $5.5 \times 10^{-12}$\,erg\,cm$^{-2}$\,s$^{-1}$. Thus, J1823$-$3021A\ dominates
the total $\gamma$-ray emission of the cluster. The combined emission
of all other MSPs in the cluster plus any off-pulse emission from J1823$-$3021A\, is
not detectable with present sensitivity. No other pulsars are detected
in a pulsation search either.
Under the assumption that the $\gamma$-ray emission originates from NGC 6624,
\cite{tkh+11} estimated the total number of MSPs to be
$N_{\rm MSP} = 103^{+104}_{-46}$. Assuming an average $\gamma$-ray
luminosity for each MSP \cite{aaa+09a,aaa+10c}, similar to the
approximation made by \cite{tkh+11}, our off-pulse flux upper limit
implies that $N_{\rm MSP} < 32$. This is consistent with the
estimate $N_{\rm MSP} = 30 \pm 15$ derived from the correlation
between $\gamma$-ray luminosity and encounter rate
\cite{aaa+10c}. Clearly, the MSP number estimate of \cite{tkh+11}
is skewed by the presence of a single bright pulsar contributing
disproportionately to its emission \cite{fg00}.
The off-pulse emission limits can also be used to constrain alternative
models for the $\gamma$-ray emission from globular
clusters, like those invoking inverse Compton (IC) radiation \cite{bs07,ccd+10}.
The spin period of J1823$-$3021A, 5.44 ms, is typical of MSPs.
However, its rate of change
$\dot{P}_\mathrm{obs} = +3.38 \times 10^{-18}$\,s\,s$^{-1}$ is one to two
orders of magnitude larger than for other MSPs with the exception of
J1824$-$2452A, a pulsar in the GC M28 \cite{lbm+87} that has a
similarly large $\dot{P}_\mathrm{obs}$ \cite{fbtg88}. A possible explanation
is that $\dot{P}_\mathrm{obs}$ is due mostly to the changing Doppler shift
caused by the pulsar's acceleration in the gravitational field of the
cluster along the line of sight ($a_l$):
\begin{equation}\label{eq:period}
\label{eq:acc}
\left( \frac{\dot{P}_{\rm obs}}{P} \right) =
\left( \frac{\dot{P}}{P} \right) + \frac{a_l}{c}.
\end{equation}
If the globular cluster has a reliable mass model, we could use it to
estimate lower and upper limits for $a_l$ and estimate upper and lower
limits for $\dot{P}$ \cite{fck+03}. For
NGC~6624 the collapsed nature of its core precludes the
derivation of a reliable mass model. Furthermore, radio
timing (Table S1) shows that J1823$-$3021A\ is only $0\farcs4 \pm 0\farcs1$ (a
projected distance of $0.018\pm0.004$ pc) from the center of the
cluster \cite{gra+10}, where the values of $a_l$ can be
largest. For this reason, it has been suggested \cite{bbl+94}
that J1823$-$3021A\, is a ``normal'' MSP (i.e., with small $\dot{P}$); its
large $\dot{P}_\mathrm{obs}$ being due to its acceleration
in the cluster. This conclusion was apparently strengthened by the
detection of a second derivative of the spin period
$\ddot{P} = -1.7 \times 10^{-29}$\,s$^{-1}$ \cite{hlk+04}.
This could originate in a time variation
of $a_l$ resulting from interaction with a nearby object \cite{phi92}.
If sustained it would reverse the sign of
$\dot{P}_{\rm obs}$ in $\sim$\,6000 years; suggesting again
that the large $\dot{P}_\mathrm{obs}$ is not only due to
dynamical effects, but is possibly a transient feature.
However, the total observed $\gamma$-ray emission $L_{\gamma}$
must represent a fraction $\eta < 1$ of the available
rotational energy loss, $\dot{E} = 4 \pi^2 I \dot{P}/ P^3$, where
$I$ is the pulsar's moment of inertia. Although $I$ depends on
the unknown mass of the pulsar and the unknown equation of state
for dense matter, the standard asumption $I = 10^{45} \rm \, g\, cm^{2}$
is a reasonable value for a 1.4-$M_{\odot}$ (mass of the Sun)
neutron star. This implies
$\dot{P} > 3.4 \times 10^{-19}\,(f_{\Omega}/0.9)(I/10^{45}\rm g\, cm^2)^{-1}$\,s\,s$^{-1}$. Thus
even an unrealistic $\gamma$-ray efficiency $\eta = 1$ would imply
that $\dot{P}$ is already $\sim$10\% of $\dot{P}_{\rm obs}$.
If we assume instead $\dot{P} \simeq \dot{P}_\mathrm{obs}$, then
$\dot{E} = 8.3\,\times\,10^{35}$\,erg\,s$^{-1}$\ and $\eta = 0.1 \times
(f_{\Omega}/0.9)(I/10^{45}\rm g\,cm^2)^{-1}$.
Comparison with the observed $\gamma$-ray
efficiencies of other MSPs \cite{LAT_MSPs,aaa+10} shows this
to be a more reasonable range of values; $\eta \sim 0.1$ also
represents the upper limit derived for the average efficiency of
MSPs in 47 Tucanae \cite{aaa+09a}. Therefore, our $\gamma$-ray
detection of J1823$-$3021A\, indicates that it is unusually energetic and
that most of $\dot{P}_\mathrm{obs}$ is due to
its intrinsic spin-down.
The pulsar has other features that suggest it is indeed unusually energetic:
Its alignment of radio and $\gamma$-ray profiles has previously
only been observed for the Crab pulsar \cite{aaa+10a} and three fast,
energetic MSPs: J1939+2134 (the first MSP to be discovered),
J1959+2048 \cite{gcj+11} and J0034$-$0534 \cite{aaa+10b}. Like some of
these energetic pulsars and PSR~J1824$-$2452A, J1823$-$3021A\, emits giant radio
pulses \cite{kni07} and has a high 400 MHz radio luminosity of
$L_{400}\,\simeq\, 1.1$ Jy kpc$^2$ \cite{bbl+94}, the third highest among
known MSPs. However the correlation between $\dot{E}$ and radio
luminosity is far from perfect given the uncertainties in the distance
estimates, moment of inertia, beaming effects and
possibly intrinsic variations of the emission efficiencies.
Finally, J1939+2134 also has a large $\ddot{P}$ \cite{cbl+95},
which is thought to be caused by timing noise (TN),
which scales roughly with $P^{-1.1}\dot{P}$
\cite{sc10}. In the case of J1823$-$3021A, if $\dot{P} \simeq \dot{P}_\mathrm{obs}$,
then TN should be one order of magnitude larger than for J1939+2134;
instead its $\ddot{P}$ is $\sim 1.5 \times 10^2$
larger than that of J1939+2134. This is possible given the observed
scatter around the TN scaling law. Thus
TN might account for the $\ddot{P}$ of J1823$-$3021A, but this is far more
likely if $\dot{P} \simeq \dot{P}_\mathrm{obs}$.
If $\dot{P} \simeq \dot{P}_\mathrm{obs}$,
we can estimate the strength of its surface dipole magnetic
field: $B_0 = 3.2 \times 10^{19} {\rm G} \sqrt{\dot{P} P (I / 10^{45}
{\rm\,g\,cm^2})} (R/ 10{\rm\,km})^{-3} \simeq 4.3 \times
10^9$ G\cite{lk04} [where $R$ is the neutron star (NS) radius, generally
assumed to be 10 km]. MSPs are thought to start
as normal NSs with $B_0 \sim 10^{11-13}$\,G which are
then spun up by the accretion of matter and angular momentum
from a companion star. This process is thought to decrease their magnetic
field to $B_0 \sim 10^{7-9}$\,G; but the exact mechanism responsible
for this is currently not well understood. Our value of
$B_0$ shows that for J1823$-$3021A\, this decrease
was not as pronounced as for other MSPs.
As accretion spins up the NS, it eventually reaches an equilibrium
spin period \cite{acrs82} given by:
\begin{equation}
\label{eq:pmin}
P_{\rm init} = 2.4 {\rm ms} \left( \frac{B_0}{10^9 {\rm G}} \right)^{6/7}
\left( \frac{M}{M_{\odot}} \right)^{-5/7}
\left( \frac{R}{10^4 \rm m} \right)^{18/7}
\left( \frac{\dot{M}}{\dot{M}_{\rm Edd}} \right)^{-3/7},
\end{equation}
where $M$ is the NS mass, $\dot{M}$ is the accretion rate
and $\dot{M}_{\rm Edd}$ is the maximum possible stable accretion rate
for a spherical configuration (known as the Eddington rate). Beyond
this, the pressure of accretion-related radiation starts preventing
further accretion. After accretion ceases, the newly formed radio MSP
will have $P_{\rm init}$ as its initial spin period.
Assuming $\dot{M} = \dot{M}_{\rm Edd}$,
$M = 1.4\,M_{\odot}$ and $R =$10~km (as in our estimates of
$B_0$), we obtain $P_{\rm init} = 1.9\,{\rm ms} ( B_0 / 10^9 {\rm
G})^{6/7}$. For the value of $B_0$ calculated above, we get $P_{\rm
init} = 6.6\,$ms; that is, even if accretion had proceeded at
the Eddington rate, the pulsar would not have been
spun up to its present spin frequency. This is also the case for
the other such ``anomalous'' MSP,
J1824$-$2452A \cite{fbtg88}; for all others we have $P > P_{\rm init}$.
A possible explanation is that for these two objects $M$ and $I$
do not correspond to the assumptions above. If, for example, $\eta\,=\,0.15$,
$M\,=\,1.8\,M_{\odot}$ and
$I\,=\,1.8 \times 10^{45} \rm g\,cm^2$ \cite{wkl08} we obtain
$B_0 = 3.6 \times 10^9$ G and
$P_{\rm init}\,=\,4.7\,{\rm ms}$. A second possibility, suggested by
eq.~\ref{eq:pmin}, is super-Eddington accretion (more precisely,
$\dot{M} > 1.6\,\dot{M}_{\rm Edd}$); this can happen for non-spherical mass
accretion. A third possibility is that the value of $B_0$ was smaller
during accretion (resulting in a smaller $P_{\rm init}$), and that
$B_0$ has increased since then. This has been observed for some normal pulsars
\cite{elk+11}; however there is no evidence of such behavior for any
other MSPs.
In any case, the conclusion that
$\dot{P} \simeq \dot{P}_\mathrm{obs}$ implies a characteristic age
$\tau_c = P / (2 \dot{P}) = 25$ million years. This is likely an
over-estimate of the true age of the pulsar, particularly given that
$P_{\rm init}$ is likely to be similar to $P$. Thus J1823$-$3021A\ is likely to
be the youngest MSP ever detected; only J1824$-$2452A might have a
comparable age. Because of their large
$\dot{P}$s both objects will be observable as MSPs for a time that
is $\sim 10^2$ shorter than the $\sim 100$ ``normal'' radio-bright
MSPs known in GCs. Statistically, this suggests that, at least in
GCs, anomalous high $B$-field
MSPs like J1823$-$3021A\, and J1824$-$2452A are forming at rates
comparable to those of the more ``normal'', radio-bright MSPs.
|
\section{Introduction}
In principle we know how to find the electronic structure and binding energies of atom: solve
the many-body time-independent Schr\"odinger equation for $N$ electrons
about a nucleus with charge $Z$:
\begin{equation}
\hat{H}\Psi = \left (\sum_{i = 1}^N -\frac{\hbar^2 }{2m} \nabla^2_i- \frac{Z e^2}{r_i}
+ \sum_{i > j} \frac{e^2}{\left|\vec{r}_i - \vec{r}_j \right|} \right)
\Psi = E \Psi
\label{Hcoord}
\end{equation}
or variants thereof including relativistic corrections. In practice one cannot solve
this exactly, or even fully numerically much beyond helium.
So one turns to approximations. But for any approximation there are errors intrinsic to
the approximation method itself, and errors due to the input, such as truncations in
the space or assumptions in the Hamiltonian. Surprisingly, rather little effort has been
put into disentangling these two sources of error.
In this paper we benchmark three approximations against an `exact' calculation with a fixed space and fixed nonrelativistic Hamiltonian. For our exact calculation we use
full configuration-interaction ( full CI); for a chosen single-particle basis, we can find numerically
converged solutions for the ground state (as well as some number of excited states).
The full CI spaces are large, up to dimension $10^8$ in the uncoupled (`M-scheme') basis,
or $2 \times 10^6$ in the coupled basis of configuration state functions (CSFs);
see below for further discussion.
We also have a suite of codes that carry out Hartree-Fock, angular-momentum projected Hartree-Fock, and
random phase approximations, using the identical input
as CI.
Thus our goal is to carefully document errors due the approximations themselves (and not due to
the single-particle space, relativistic corrections or lack thereof, etc.); in this we are
extending prior work benchmarking approximations for the nuclear many-body problem \cite{SJ02,SJ03,PHF}.
In the next section we define our many-body methods and discuss briefly computational
issues. In Section 3 we give our results for several different spaces for second row atoms, from Li to
Ne, comparing not only the ground state binding energies but also the experimentally more relevant
ionization potentials and electron affinities. Because full CI is very time-consuming and our
approximations computationally cheap, in Section 4 we consider the possibility of using
approximations for fast optimization of the the single-particle basis, and of the considered
methods find RPA the most accurate.
\section{Many-body methods and inputs}
We work entirely in occupation or configuration space, meaning
we start with some set of orthonormal single-particle states,
$\{ \phi_a(\vec{r}) \}$, with good angular momentum and parity, and introduce
in second quantization \cite{LM85, BG77} fermion creation and annihilation
operators $\hat{a}^\dagger, \hat{a}$ that create or remove a particle from
the states $ \phi(\vec{r})$. For definitiveness we have $N_s$ single-particle states.
The Hamiltonian is then \cite{LM85,BG77,ring}
\begin{equation}
\hat{H}=\sum_{ab} H^1_{ab} \hat{a}^\dagger_a \hat{a}_b
+ \frac{1}{4} \sum_{abcd} V_{ab,cd}\hat{a}^\dagger_a \hat{a}^\dagger_b \hat{a}_d \hat{a}_c
\label{H2q}
\end{equation}
The one-body and two-body matrix elements, $H^{1}_{ab}$ (which includes both kinetic
energy and the nuclear central potential) and $V_{ab,cd}$ respectively,
are appropriate integrals over the Hamiltonian and the single-particle states.
Because the Hamiltonian is an angular momentum scalar, we generate the matrix elements
in coupled form. Given the single-particle space, the matrix elements are computed externally to
all our codes and read in via a file. Any radial form of the single-particle states and any form of the
interaction, including non-local interactions, are allowed and make no practical difference to our
many-body codes. Because we work in second quantization, the matrix elements are antisymmeterized and
we do not separate direct from exchange terms.
For this paper we restrict the Hamiltonian to the simple nonrelativistic Coulomb interaction (\ref{Hcoord}).
We build our single-particle radial wavefunctions from Slater-type orbitals (STOs), that is,
functions of the form $r^n \exp(-\beta r)$, which allows us to compare to other work as well as making
the integrals analytic. We orthonormalize our single-particle basis so that the one-body part of
(\ref{H2q}) is diagonal.
For our specific choices of parameters for our basis, we use the CVB$n$ bases sets \cite{Ema03},
described below.
Finally, we describe the atom as a two-species systems, with spin-up and spin-down electron occupying
distinct orbits. We fix $M_s = (n_\uparrow - n_\downarrow)/2$ in all calculations, as
we have no spin-orbit coupling.
These methods were chosen because we have existing codes for them, and thus leave for future work
other methods such as the multi-configurational Hartree-Fock approximation \cite{FFBJ97,Cook98}
and many-body
perturbation theory \cite{LM85,Je07,Cook98}
\subsection{Configuration-interaction calculations}
For full configuration-interaction (full CI) calculations \cite{BG77,Lo55CI,CIV3,GRASP92, SS99,Cook98,Je07,BW88,Sh98},
we use the BIGSTICK code \cite{BIGSTICK}, developed originally for
nuclear CI calculations but which can be applied as well to atomic calculations. The only assumption is that the
single-particle basis be states of good angular momentum and parity.
BIGSTICK creates a many-body basis of Slater determinants $\{ | \alpha \rangle \}$, constructed
from the single-particle basis:
\begin{equation}
| \alpha \rangle = \prod_{i = 1}^N \hat{a}^\dagger_i | 0 \rangle,
\end{equation}
where $N$ is the number of particles (occupied states).
Furthermore, because we have no spin-orbit coupling (a generalization we leave to future work), we fix
the number of spin-up and spin-down electrons and thus fix $M_s = (n_\uparrow -n_\downarrow)/2$.
Because single-particle states each have good $l_z$,
it is almost trivial to generate many-body basis states with fixed total $M = L_z$; in nuclear
CI calculations this is called an $M$-scheme basis. And because the Hamiltonian is an angular momentum scalar,
the final eigenstates thus will automatically have good total $L$ (and $S$). Although we have the capability
for various truncations on the many-body basis, such as allowing only single- or double-electron excitations,
we make no such truncations here as they have no correspondence for our approximation methods.
Most atomic CI programs such as CIV3 \cite{CIV3} and
GRASP92 \cite{GRASP92} use a basis of configuration state functions (CSFs) coupled up to good
$L$ and $S$ (or total $J$, hence in nuclear structure physics this is called a $J$-scheme basis).
Such a basis is smaller in dimension, by a factor of 10 to 100, but there are always trade-offs: the basis states are linear
combinations of $M$-scheme Slater determinants and hence the many-body Hamiltonian matrix elements are
more complicated to calculate. And although the dimensions of an uncoupled $M$-scheme basis are larger,
the Hamiltonian matrix is much sparser. Furthermore the simplicity of the uncoupled basis allows us to avoid
storing the many-body matrix element, and instead recompute them
efficiently on the fly \cite{ANTOINE}, which further reduces
the storage of the Hamiltonian by a factor of 10 to 100.
We can go up to uncoupled basis dimensions of $10^8$ on a desktop machine (and similar programs on parallel
computers reach up to $10^{10}$), as listed in Table II, while the dimension of the largest equivalent
coupled CSF ground state dimension is about $ 10^7$, as listed in Table III.
Given these arguments, it is interesting to compare basis sizes. These are not always given in
the literature, and many `large-scale' calculations rely upon careful pruning
of the configuration basis \cite{Correge04,Gupta05,Deb09}; nevertheless, for comparison
purposes a large scale relativistic CI calculation \cite{Chen01}
used up to 197,426 basis CSFs, while a more recent nonrelativistic calculation \cite{Bro10} used up to
901,816 CSFs. Quantum chemists surpassed the one-billion determinant limit some time ago \cite{Sh98,OJS90}.
In a given uncoupled basis, BIGSTICK computes the many-body Hamiltonian matrix elements
$\langle \alpha | \hat{H} | \beta \rangle$ and finds the low-lying eigenstates, including the ground state, using the Lanczos algorithm \cite{Lanczos} (the Davidson-Liu algorithm \cite{Dav93,LSAS01} is efficient for the
diagonal-dominated atomic problem but not the more general many-body problem such as nuclear CI; furthermore recomputing the matrix elements on the fly is better suited for the Lanczos algorithm).
Convergence of, say, five states take a few hundred Lanczos iterations;
on a desktop machine with eight OpenMP threads
we can accomplish this for an $M$-scheme basis in a few hours or at most a few days.
With the single-particle space and Hamiltonian matrix elements fixed, the resulting eigenenergies for the
full-space CI calculation are numerically exact. All other calculations described in this paper are
numerical approximations to these results.
CI calculations have been previously carried out for second row elements, see for example
nonrelativistic \cite{Weiss61,Bu68,Saski74,Bunge76,Silver79,Jitrik97,Bunge06} and relativistic calculations \cite{Chen01,Almora10} , though
most of these were not full CI and thus were in much smaller dimensions.
\subsection{Hartree-Fock approximation}
The Hartree-Fock approximation (HF) is a variational method using a single Slater determinant \cite{Lo55HF,FFBJ97,Cook98,Je07}. In occupation space \cite{ring} one applies
a unitary transformation among the single-particle states:
\begin{equation}
\hat{c}_i^\dagger = \sum_{a=1}^{N_s} U_{ia} \hat{a}_a^\dagger, \label{Usp}
\end{equation}
and creates the trial Slater determinant
\begin{equation}
| \Psi_T \rangle =\prod_{i =1}^N \hat{c}^\dagger_i | 0 \rangle.
\end{equation}
Then one finds a Slater determinant that minimizes the energy, that is, that minimizes
\begin{equation}
E = \frac{ \langle \Psi_T | \hat{H} | \Psi_T \rangle } { \langle \Psi_T | \Psi_T \rangle }
\end{equation}
Because our input matrix elements are antisymmeterized, we fully include both direct and exchange
terms. Working in occupation space the exchange term causes us no difficulty (or, to put it another way,
any difficulty is off-shored into calculation of the antisymmeterized integrals).
Many Hartree-Fock calculations enforce good angular momentum, for closed-shell systems or
closed-shell plus or minus one particle. Our HF code allows the Slater determinant to break rotational
invariance, even for closed-shell systems; such Slater determinants are `deformed.'
Experience in nuclear physics suggest this to be a fruitful
path. In a few cases, when possible, we will compare and contrast spherical and deformed HF solutions.
In order to minimize, we use the standard Hartree-Fock equations, which consist of
iteratively solving
\begin{equation}
\mathbf{h}\vec{u}_i = \epsilon_i \vec{u}_i, \label{HF}
\end{equation}
where the Hartree-Fock effective one-body Hamiltonian is
\begin{equation}
h_{ab} = T_{ab} + \sum_{cd} V_{ac,bd} \rho_{dc}
\end{equation}
and
\begin{equation}
\rho_{dc} = \sum_{i = 1}^N U_{ic} U_{id}.
\end{equation}
Internally we use an $N_s \times N$ rectangular matrix $\mathbf{\Psi}$, with the $N$
columns representing the transformed occupied states $\hat{c}^\dagger_i$. In that
case $\mathbf{\rho} = \mathbf{\Psi \Psi}^\dagger$.
(Because we leave out spin-orbit coupling and conserve $n_\uparrow$ and $n_\downarrow$,
we use separate Slater determinants for spin-up and spin-down electons; this is a
straightforward generalization. )
The Hartree-Fock energy is now
\begin{equation}
E_\mathrm{HF} = \sum_{i=1}^N \left( \epsilon_i -\frac{1}{2} \sum_{cd}
V_{ic,id} \rho_{dc} \right),
\end{equation}
where we have to subtract off the potential to keep from double-counting in the
single-particle energies.
\subsection{Projected Hartree-Fock approximation}
If one does not have a closed shell, a single Slater determinant will not in general
have good angular momentum. Conversely, an open-shell state with good angular momentum,
which can be written as a sum of Slater determinants, will have important correlations that
lower the energy.
One can start with single-particle states which have good angular momentum and then construct
many-body states with good total angular momentum; these are called configuration state functions (CSFs),
and have been applied to second-row elements, e.g. \cite{ Lask75,Lunell68}.
We take a different approach, however. Our Hartree-Fock code allows for arbitrary deformation:
the single-particle states generated by the self-consistent field do not have good angular momentum.
The Slater determinant built from these states is called a `deformed' state.
In order to project out good angular momentum, we introduce the
standard projection operator \cite{La80,ring}
\begin{equation}
\hat{P}^J_{MM^\prime} =
\int d\Omega {\cal D}^{(J)*}_{MM^\prime} (\Omega) \hat{R}(\Omega),
\label{jproj0}
\end{equation}
where $\Omega =$ the Euler angles $ \alpha,\beta,\gamma$ and $d\Omega = d\alpha \sin \beta d\beta d\gamma$,
${\cal D}^{(J)*}_{MM^\prime}$ is the Wigner $D$-matrix \cite{Edmonds}, and
$\hat{R}(\Omega)$ is the rotation operator.
To compute
$\langle \Psi | \hat{R}(\Omega) | \psi \rangle$ and
$\langle \Psi | \hat{H} \hat{R}(\Omega) | \psi \rangle$, we use the
matrix representation
\begin{equation}
\hat{R}| \Psi \rangle \rightarrow \mathbf{R} \mathbf{\Psi} = \tilde{\mathbf{\Psi}}
\end{equation}
where $\mathbf{\Psi}$ and $\tilde{\mathbf{\Psi}}$ are $N_s \times N$ matrices
representing our original and rotated Slater determinants, respectively, and
$\mathbf{R}$ is a square $N_s \times N_s$ matrix; in our single-particle
basis the matrix elements are straightforward:
\begin{equation}
R_{ab} = \langle j_a m_a | \hat{R}(\alpha, \beta, \gamma) | j_b m_m \rangle
= \delta_{j_a j_b} {\cal D}^{j_a}_{m_a, m_b}(\alpha, \beta, \gamma),
\end{equation}
where ${\cal D}^{j}_{m^\prime, m}$ is the Wigner $D$-matrix \cite{Edmonds}.
It is useful to note that the $j,m,m^\prime$ for the rotational matrix $\hat{R}$ are those
of the single-particle space, while the $J,M, M^\prime$ for the projection operator
(\ref{jproj0}) are those of the many-body space.
Matrix elements between two oblique Slater determinants is
relatively straightforward \cite{Lang93}. First,
\begin{equation}
\langle \Psi | \hat{R}|{\Psi} \rangle =
\langle \Psi | \tilde{\Psi} \rangle = \det \Psi^\dagger \tilde{\Psi}.
\end{equation}
Calculation of the Hamiltonian (and other) expectation value requires
the density matrix,
\begin{equation}
\rho^\prime_{ab} = \frac{ \langle \Psi | \hat{\phi}^\dagger_a
\hat{\phi}_b|\tilde{\Psi} \rangle }
{ \langle \Psi | \tilde{\Psi} \rangle}
= \left [
\tilde{\mathbf{\Psi}} \left( \mathbf{\Psi}^\dagger \tilde{\mathbf{\Psi}}\right )^{-1}
\mathbf{\Psi}^\dagger \right ]_{ba}
\end{equation}
Then
\begin{equation}
\langle \Psi | \hat{H} | \tilde{\Psi} \rangle
= \sum_{ab} H^1_{ab} \rho^\prime_{ab}
+ \frac{1}{2}\sum_{abcd}V_{ab,cd}( \rho^\prime_{ac} \rho^\prime_{bd} - \rho^\prime_{ad} \rho^\prime_{bc} )
\end{equation}
Accounting for two species is straightforward.
Note that even if our original Slater determinants were real, by rotating over all
angle they can become complex.
We only project out good orbital angular momentum $L$. We could in principle project out
exact spin $S$; we leave such a modification for
future work. Because we leave off spin-orbit coupling, however, we automatically get states
with good $S$ already.
We now can compute the Hamiltonian and ``norm'' matrix elements,
\begin{equation}
h^J_{M,M^\prime} \equiv \langle \Psi|
\hat{H} \hat{P}^J_{M,M^{\prime}} | \Psi \rangle , \,\,\,
n^J_{M,M^\prime} \equiv \langle \Psi|
\hat{P}^J_{M,M^{\prime}} | \Psi \rangle,
\end{equation}
which are both Hermitian.
Then we solve the
generalized eigenvalue equation
\begin{equation}
\sum_{M^\prime} h^J_{M,M^\prime} g^J_{M^\prime} = E^J_{PHF} \sum_{M^\prime} n^J_{M,M^\prime} g^J_{M^\prime}.
\end{equation}
where we allow $g_M$ to be complex. Although the deformed
Hartree-Fock state will have an orientation, the final result will be independent of
orientation; we confirmed this by arbitrarily rotating our HF state.
Both the Hartree-Fock and the projected Hartree-Fock energies will be variational
upper bounds to the exact CI energy, with the projected Hartree-Fock energy lower than
(or equal to, in the HF state has good rotational symmetry) $E_{HF}$.
\subsection{Random phase approximation}
Once one has a Hartree-Fock calculation, it is straightforward to go on to the random
phase approximation, using the matrix formulation \cite{ring,SJ02}:
\begin{equation}
\left(
\begin{array}{cc}
\mathbf{A} & \mathbf{B} \\
-\mathbf{B}^* & -\mathbf{A}^*
\end{array}
\right ) \left( \begin{array}{c} \vec{X}_\lambda \\
\vec{Y}_\lambda
\end{array} \right )
= \hbar \Omega_\lambda \left( \begin{array}{c} \vec{X}_\lambda \\
\vec{Y}_\lambda
\end{array} \right )
\end{equation}
where
\begin{equation}
A_{nj,mi} \equiv \left \langle \left [\left [ \hat{c}^\dagger_j \hat{c}_n, \hat{H} \right ], \hat{c}^\dagger_m
\hat{c}_i \right ] \right \rangle
\end{equation}
and
\begin{equation}
B_{nj,mi} \equiv \left \langle \left [ \hat{c}^\dagger_i
\hat{c}_m ,\left [ \hat{c}^\dagger_j \hat{c}_n, \hat{H} \right ] \right ] \right \rangle.
\end{equation}
Here $i,j$ label occupied single-particle levels in the Hartree-Fock state,
while $m,n$ label unoccupied levels.
From this we can compute the RPA correlation energy, which is a correction to the
HF energy. When working with a rotationally invariant Hamiltonian, as we do here,
it is possible to have a HF state which breaks rotational symmetry; this means one
could rotate the HF state in any direction and not change the HF energy. In RPA
this shows up as a zero-energy mode, which must be dealt with carefully
when computing the binding energy \cite{ring,SJ02}. The final result is
\begin{equation}
E_\mathrm{RPA} = E_{HF} -\frac{1}{2}{\rm Tr}(A) + \frac{1}{2}\sum_{i}\hbar \Omega_i.
\end{equation}
The RPA energy, while lower than the HF energy, is not variational. In previous
calculations for nuclei in valence spaces \cite{SJ02} RPA both under- and overestimated
the binding energy; in our chosen atomic cases below, RPA consistently overestimates the
binding energy. Because we start from a deformed HF state, our RPA calculations do
not have good angular momentum. (RPA is often said to `restore' rotational symmetry,
but this is not restoration of good angular momentum quantum numbers. Rather,
the Hartree-Fock energy is invariant under rotation, and this invariance shows up as
a $\Omega=0$ mode that appears in RPA but not in the simpler Tamm-Dancoff approximation
or TDA \cite{ring}. In this sense RPA `restores' rotational invariance that was broken in TDA.)
\subsection{Input spaces and interaction}
Our codes can use (up to memory limitations, of course) any finite set of
single-particle states which have good angular momentum. For convenience we
use the CVB1, CVB2, and CVB3 Slater-type orbitals (STOs) and parameter values of Ema
\textit{et al.} \cite{Ema03}. These are radial wavefunctions of the
form $r^n \exp(-\beta r)$, with the set of STOs and the values of $\beta$ optimized
for Li through Ne. The single-particle spaces ranged from 8 orbits
(for CVB1-Li) up to 22 (for CVB3-Ne).
\begin{table}[h]
\centering
\caption[Basis Set Composition]{Basis Set Composition \label{table:basislist}}
\begin{tabular}{|c|ccccc|cccc|}
\hline
\multicolumn{2}{|c}{~~}&~~&\multicolumn{3}{c}{Li and Be}&~~&\multicolumn{3}{c|}{B to Ne}\\
Orbit&&&CVB1&CVB2&CVB3&&CVB1&CVB2&CVB3\\
\hline
s&&&6&7&8&&6&7&8\\
p&&&2&3&4&&4&5&6\\
d&&&0&2&3&&1&3&4\\
f&&&0&0&2&&0&1&3\\
g&&&0&0&0&&0&0&1\\
\hline
Total&&&8&12&17&&11&16&22\\
\hline
\end{tabular}
\end{table}
To construct our many-electron basis states, we fixed $n_\uparrow$ spin-up
electrons and $n_\downarrow$ spin-down electrons, with $n_\uparrow = n_\downarrow$
or $=n_\downarrow+1$, which allowed us to regain states of all possible total $S$. We could and did choose
$n_\uparrow + n_\downarrow \neq Z$ in order to compute ionization potentials and
electron affinities. We also fixed total $M = L_z = 0$, which allowed us to
regain states of all possible total $L$. We otherwise had no constraints, e.g., we did not
restrict ourselves to single, double, or otherwise excitations; although our
CI code can make such restrictions, the HF, PHF, and RPA cannot. Dimensions
of the uncoupled many-body CI spaces we used are in Table \ref{table:cispace}, while the
equivalent dimensions for coupled (CSF) bases for the ground states are in Table
\ref{table:cispacecoupled}, where we indicated the ground state quantum numbers
by $^{2S+1}(L)^\pi$.
\begin{table}[h!]
\centering
\caption[Size of the CI Space]{Dimensions of the uncoupled ($M$-scheme) CI spaces \label{table:cispace}}
\begin{tabular}{|l|rrrr|}
\hline
Atom &$~$& \multicolumn{3}{c|}{Basis Set}\\
&&\multicolumn{1}{c}{CVB1} & \multicolumn{1}{c}{CVB2} & \multicolumn{1}{c|}{CVB3} \\
\hline
Li && $192$ & $1067$ & $4958$ \\
Be && $866$ & $10983$ & $100842$ \\
B && $46116$ & $909534$ & $1.1\times10^7$ \\
C && $300309$ & $1.1\times10^7$ & $2.5 \times 10^8$ \\
N && $1.4\times10^6$ & $1.1\times10^8$ & \multicolumn{1}{c|}{--}\\
O && $6.6\times10^6$ &\multicolumn{1}{c}{--} & \multicolumn{1}{c|}{--}\\
F && $2.4\times10^7$ &\multicolumn{1}{c}{--} & \multicolumn{1}{c|}{--} \\
Ne && $8.7\times10^7$ &\multicolumn{1}{c}{--} & \multicolumn{1}{c|}{--} \\
\hline
\end{tabular}
\end{table}
\begin{table}[h!]
\centering
\caption[Dimensions of the Coupled CI Space]{Size of the equivalent coupled (CSF or $J$-scheme) CI spaces \label{table:cispacecoupled}}
\begin{tabular}{|l|c|rrrr|}
\hline
Atom & $^{2S+1}(L)^\pi$ &$~$& \multicolumn{3}{c|}{Basis Set}\\
& (g.s.)&& \multicolumn{1}{c}{CVB1} & \multicolumn{1}{c}{CVB2} & \multicolumn{1}{c|}{CVB3} \\
\hline
Li & $^2 S^+$ && $94$ & $223$ & $516$ \\
Be & $^1 S^+$ && $190$ & $775$ & $3102$ \\
B & $^2 P^-$ && $7752$ & $82429$ & $618245$ \\
C & $^3 P^+$ && $35126$ & $770145$ & $1.1 \times 10^7$ \\
N & $^4 S^-$ && $53710 $ & $2.1\times 10^6 $ & \multicolumn{1}{c|}{--}\\
O & $^3 P^+$ && $604146$ &\multicolumn{1}{c}{--} & \multicolumn{1}{c|}{--}\\
F & $^2 P^-$ && $1.8 \times 10^6$ &\multicolumn{1}{c}{--} & \multicolumn{1}{c|}{--} \\
Ne & $^1 S^+$ && $1.3 \times 10^6$ &\multicolumn{1}{c}{--} & \multicolumn{1}{c|}{--} \\
\hline
\end{tabular}
\end{table}
To construct our interaction file, we first orthornormalized the STO basis.
We did this by diagonalizing the one-body part of the Hamiltonian (and so this
depended on the $Z$ charge of the nucleus); then the one-body Hamiltonian is
diagonal and expressed in terms of single-particle energies. The two-body matrix
elements were first computed with the STOs, which can be done semi-analytically, and
then transformed to the orthonormalized basis. Our matrix elements and resulting
atomic energies for small cases were validated against an prior, independent code (M. Bromley,
private communication).
\section{Results}
The following sections give descriptions of the results obtained as well as tables and example figures. The computed ground state energies for HF, PHF, and RPA are compared to CI. The ionization potentials and electron affinities are compared to experimental values
\cite{lide}.
\subsection{Ground State Energy}
The computed ground state energies for the neutral atoms studied are in Table \ref{tab:gs}. CI is exact for purposes of comparison. As required by variational theory, Hartree-Fock gives an upper bound to the ground state energy and projected Hartree-Fock improves upon it. The random phase approximation computes lower energies than CI in all of the calculations we performed. There are no CI calculations for several the largest basis sets in combination with the largest atoms because the computational requirements exceeded our available resources; keep in mind that our `approximate' results were computed in
full CI spaces, as we allow arbitrary deformation of the single-particle states.
Figure \ref{fig:carbongs} shows the results for carbon in graphical form; the results for all other
atoms are qualitatively similar, although the error for all approximations increases as we add more
electrons. For example, the error in RPA is only a few millihartrees for Li, while it is 100 millihartrees
for Ne.
Note there is no difference between the HF and PHF results for Li and for Ne; this is because
the HF states for neutral
Li and Ne are already in states of good total L; projection cannot change this.
\begin{table}
\caption[Ground State Energy]{Ground State Energy (hartree)\label{tab:gs}}
\begin{tabular}{lr|rrrrrrr}
\hline
\multicolumn{1}{c}{Atom} & \multicolumn{1}{c}{space} & \multicolumn{1}{c}{HF} & & \multicolumn{1}{c}{PHF} & &\multicolumn{1}{c}{RPA} & & \multicolumn{1}{c}{CI} \\
\hline
& CVB1 &-7.4327 & &-7.4327& &-7.4720 & &-7.4700 \\
Li & CVB2 &-7.4327 & &-7.4327& &-7.4751 & & -7.4725 \\
& CVB3 &-7.4327 & &-7.4327& &-7.4774 & &-7.4741 \\
\hline
& CVB1 &-14.5732& &-14.5758& &-14.7491& &-14.6558 \\
Be &CVB2 &-14.5733& &-14.5778& &-14.7271& &-14.6589 \\
& CVB3 &-14.5733& &-14.5780& &-14.7315& &-14.6615 \\
\hline
& CVB1 &-24.5239& &\multicolumn{1}{c}{-24.5352}& &-24.6852 & &-24.6391 \\
B & CVB2 &-24.5330& &-24.5353& &-24.6965& &-24.6458 \\
& CVB3 &-24.5331& &-24.5355& &-24.7009& &-24.6482 \\
\hline
& CVB1 &-37.6700 & &-37.7052 & &-37.8453& &-37.8202 \\
C & CVB2 & -37.6715 & &-37.7061& &-37.8661& &-37.8326 \\
& CVB3 &-37.6717& &-37.7072& &-37.8753& &-37.8378 \\
\hline
& CVB1 &-54.3443& &-54.4221& &-54.5671& &-54.5517 \\
N & CVB2 &-54.3445& &-54.4211& &-54.5971& &-54.5699 \\
& CVB3 &-54.3473& &-54.4236& &-54.6121& &\multicolumn{1}{c}{--} \\
\hline
& CVB1 &-74.7793& &-74.8267& &-75.0487& &-75.0037 \\
O & CVB2 &-74.7822& &-74.8303& &-75.0995& &\multicolumn{1}{c}{--} \\
& CVB3 &-74.7829& &-74.8309& &-75.1172& &\multicolumn{1}{c}{--} \\
\hline
& CVB1 &-99.4133& &-99.4156& &-99.7283 & &-99.6457 \\
F & CVB2 &-99.4155& &-99.4188& &-99.7943 & &\multicolumn{1}{c}{--} \\
& CVB3 &-99.4162& &-99.4195& &-99.8184 & &\multicolumn{1}{c}{--} \\
\hline
& CVB1 &-128.5455& &-128.5455& &-128.9307& &-128.8278 \\
Ne & CVB2 &-128.5462& &-128.5463& &-129.0083& &\multicolumn{1}{c}{--} \\
& CVB3 &-128.5470& &-128.5470& &-129.0399& &\multicolumn{1}{c}{--} \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\centering
\includegraphics[width=0.6\textwidth, trim=0mm 0mm 0mm -25mm]{gsC.eps}
\caption{(Color online.) Carbon ground state energy. \label{fig:carbongs}}
\end{figure}
\subsection{Ionization Potential}
Table \ref{tab:ion} shows the single ionization potentials for all methods as well as experimental values \cite{lide} for all atoms studied. Figure \ref{fig:carbonion} shows the ionization potential for carbon graphically.
Because the ionization potential (and in the next section, the electron affinity) is the difference between two ground state energies, the nice clean trends of the previous section are lost.
CI agrees well with experiment, within a few millihartree through C and less than 20 millihartree difference
through Ne. Increasing basis size brings better agreement with experiment, though of course we note
we have left out relativistic corrections.
HF consistently underestimates the ionization potentials, with PHF closer for Li through N (with
HF and PHF identical for Li as the neutral Li and Li I HF states both have good angular momenta) but
farther away for O, F, and Ne.
RPA results are the sporadic, sometimes better than HF / PHF and sometimes; again, this is because
errors in the approximation do not cancel but exacerbate. Increasing the basis size does not always
lead to better estimates; they get worse for O, F, and Ne.
\begin{table}
\caption[Ionization Potential]{Ionization Potential (hartree)\label{tab:ion}}
\begin{tabular}{lr|rrrrrrrr|r}
\hline
\multicolumn{1}{c}{Atom} & \multicolumn{1}{c}{space} & \multicolumn{1}{c}{HF} & & \multicolumn{1}{c}{PHF} & &\multicolumn{1}{c}{RPA} & & \multicolumn{1}{c}{CI} & & \multicolumn{1}{c}{Exp \cite{lide}}\\
\hline
& CVB1 &0.1963& &0.1963& &0.1972& &0.1969 & & \\
Li &CVB2 &0.1963& &\multicolumn{1}{c}{0.1963}& &0.1974& &0.1971& & 0.1982\\
& CVB3 &0.1963& &0.1963& &0.1982& &0.1978& &\\
\hline
& CVB1 &0.2957& &0.2993& &0.4324& &0.3407& & \\
Be & CVB2 &0.2958& &0.3004& &0.4072& &0.3411& & 0.3427\\
& CVB3 &0.2958& &0.3006& &0.4090& &0.3416& & \\
\hline
& CVB1 &0.2955& &\multicolumn{1}{c}{0.2978}& &0.1908& &0.3002& & \\
B & CVB2 &0.2955& &0.2978& &0.1949& &0.3030& & 0.3051\\
& CVB3 &0.2955& &0.2979& &0.1974& &0.3041& & \\
\hline
& CVB1 &0.3736& &0.4063& &0.3884& &0.4074 & & \\
C & CVB2 &0.3748& &0.4068& &0.3973& &0.4119 & & 0.4140\\
& CVB3 &0.3750& &0.4074& &0.4000& & 0.4133 & & \\
\hline
& CVB1 &0.4860& &0.5166& &0.5218& &0.5259 & & \\
N & CVB2 &0.4867& &0.5178& &0.5312& &0.5325 & & 0.5383\\
& CVB3 &0.4870& &0.5168& &0.5343& &\multicolumn{1}{c}{--} & & \\
\hline
& CVB1 &0.4857& &0.4320& &0.5201& &0.4787 & & \\
O &CVB2 &0.4852& &0.4335& &0.5350& &\multicolumn{1}{c}{--} & & 0.5007\\
& CVB3 &0.4859& &0.4340& &0.5408& &\multicolumn{1}{c}{--} & & \\
\hline
&CVB1 &0.6241& &0.5674& &0.6670& &0.6216& & \\
F &CVB2 &0.6215& &0.5652& &0.6792& &\multicolumn{1}{c}{--} & & 0.6405\\
& CVB3 &0.6220& &0.5674& &0.6858& &\multicolumn{1}{c}{--} & & \\
\hline
& CVB1 &0.7258& &0.7235& &0.8031& &0.7769 & & \\
Ne & CVB2 &0.7220& &0.7186& &0.8128& &\multicolumn{1}{c}{--} & & 0.7928\\
& CVB3 &0.7227& &0.7192& &0.8206& &\multicolumn{1}{c}{--} & & \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\centering
\includegraphics[width=0.6\textwidth, trim=0mm 0mm 0mm -25mm]{ionC.eps}
\caption{(Color online.) Carbon ionization potential. \label{fig:carbonion}}
\end{figure}
\subsection{Electron Affinity}
Not all of the second row atoms can bind an additional electron; we only present those cases which
do so experimentally: lithium, boron, carbon, oxygen and fluorine. Table \ref{tab:affinity} shows the electron affinities computed by each method and the respective experimental values \cite{lide}. Our results are comparable to other work in the field \cite{Cole82}. Figure \ref{fig:carbonaffin} shows the electron affinity of carbon in graphical form.
The quality of agreement is generally poor, even for CI. In retrospect this is not surprising:
we have already seen errors for all methods grow rapidly with the number of electrons, and given the
smallness of second row electron affinities the results are unsurprising if disappointing. Many of
the affinities even have the wrong sign. Adding an electron to the atom increases the runtimes and memory requirements over the neutral atom calculations, therefore, only the smallest basis sets for all but lithium could be used. Without the larger calculations, a clear trend is difficult to establish. It appears that the CVB1 basis set is fairly poor when computing electron affinities, sometimes giving negative results. However, based on the CVB2 calculations for boron and carbon, the CI electron affinities might converge on the experimental values as basis size increases, similar to what is seen with the ground state energy and ionization potential calculations.
All three of our approximations perform poorly, trending neither towards the CI result nor towards
experiment. HF and PHF are unable to provide consistent results across either atoms or basis sets and generally tend to be too far from experimental values to be useful; nor do they show consistent trends as the basis size is increased. RPA performs better, but our results do not suggest convergence with basis size. Because it is not clear if this is a fault of the approximation or of the basis for CI (which
again is our numerically exact benchmark) it would be good, if and when possible, to continue this
investigation with even larger bases.
\begin{table}
\caption[Electron Affinity]{Electron Affinity (hartree)\label{tab:affinity}}
\begin{tabular}{lr|rrrrrrrr|r}
\hline
\multicolumn{1}{c}{Atom} & \multicolumn{1}{c}{space} & \multicolumn{1}{c}{HF} & & \multicolumn{1}{c}{PHF} & &\multicolumn{1}{c}{RPA} & & \multicolumn{1}{c}{CI} & & \multicolumn{1}{c}{Exp \cite{lide}}\\
\hline
& CVB1 &-0.0152& &-0.0152& &0.0780& &0.0111& & \\
Li &CVB2 &0.0978 & &0.0050& &0.0804& &0.0226& &0.0227 \\
&CVB3 &-0.0020& &0.0056& &0.0768& &0.0226& & \\
\hline
& CVB1&-0.0379& &\multicolumn{1}{c}{0.0163}& &-0.0302& &-0.0141 & & \\
B & CVB2 &-0.0222& &-0.0413& &0.0009& &0.0068 & &0.0102 \\
&CVB3 &-0.0205& &-0.0004& &0.0455& &\multicolumn{1}{c}{--} & & \\
\hline
& CVB1 &-0.0126& &0.0081& &0.0186& &0.0178& & \\
C & CVB2 &0.0057 & &0.0203& &0.0591& &0.0446 & &0.0464 \\
& CVB3 &0.0060& &0.0192& &0.0616& &\multicolumn{1}{c}{--}& &\\
\hline
& CVB1 &-0.0110& &-0.0564& &0.0399& &-0.0047& &\\
O & CVB2 &0.0089& &-0.0302& &0.1092& &\multicolumn{1}{c}{--}& &0.0537 \\
& CVB3 &0.0148& &-0.0302& &0.1148& &\multicolumn{1}{c}{--}& & \\
\hline
& CVB1 &0.0133 & &0.0090 & &0.0890& &0.0600& & \\
F & CVB2&0.0437& &0.0405& &0.1706& &\multicolumn{1}{c}{--}& &0.1250 \\
&CVB3 &0.0432& &0.0399& &0.1748& &\multicolumn{1}{c}{--}& & \\
\hline
\end{tabular}
\end{table}
\begin{figure}[h]
\centering
\includegraphics[width=0.6\textwidth, trim=0mm 0mm 0mm -25mm]{affinc.eps}
\caption{(Color online.) Carbon electron affinity. There are no CI calculation for the CBV3
sets because the computational requirements exceeded our available resources. \label{fig:carbonaffin}}
\end{figure}
\subsection{Optimization of the space}
We utilized the CVB$n$ STO-basis sets , which had been arrived at
through optimization using both Hartree-Fock and CI calculations \cite{Ema03}.
But CI calculations, particularly in large
dimension spaces, are tedious and time-consuming. Furthermore, a basis optimized for Hartree-Fock calculations may not be optimized for CI calculations.
Therefore it might be useful to see how well our approximate methods can work as a
proxy for full CI when changing the basis. In other words, if we vary the basis,
does a minimum in CI have a corresponding minimum in one or more of our approximate
methods?
We carried out an example optimization by altering all of the parameter values in the STO for carbon CVB1 by a scaling factor ranging from $0.5$ to $1.5$.
Table \ref{table:carbonminima} gives the results of a quadratic fit to each computational method.
The data show that RPA has a similar minimum to CI. Figure \ref{fig:carbonopt} shows these results graphically.
\begin{table}[hbt]
\centering
\caption{Carbon Minimum Energy Scale Factor \label{table:carbonminima}}
\begin{tabular}{lc}
\hline
Method&Scale Factor\\
\hline
CI &0.981\\
RPA&0.989\\
HF &0.892\\
PHF&0.908\\
\hline
\end{tabular}
\end{table}
\begin{figure}
\centering
\includegraphics[width=0.6\textwidth, trim=0mm 0mm 0mm -25mm]{carbonopt.eps}
\caption{(Color online.) Carbon optimization. As the STO basis parameters are scaled by a scale facter between $0.5$ and $1.5$, the random phase approximation tends to track CI very well. Hartree-Fock and projected Hartree-Fock follow each other but not CI. This indicates that RPA can be useful as a proxy for CI during basis parameter optimization. \label{fig:carbonopt}}
\end{figure}
This optimization suggests that RPA is the most desirable method of those studied for basis set optimization. It is much faster than CI and clearly tracks the ground state energy as basis parameters are altered. PHF is a more useful upper bound to the ground state energy than HF, but cannot track the CI energy well enough to be used as a proxy for CI.
\section{Conclusions and Future work}
This paper is an extension of previous work comparing the random phase approximation against CI calculations
in a CI basis \cite{SJ02}, testing \textit{how good is an approximation against a
numerically exact result}. By comparing against a numerically exact result, we
eliminate any errors due to choice of model space, interaction, etc. As part of this process
we find eigenvalues of very large CI matrices.
To simplify matters we used an uncoupled basis, which allowed us to recomputed the many-body matrix
elements on the fly in a highly efficient manner.
Not surprisingly, we found in absolute error RPA gave the best approximation to the
ground state energy. RPA is not variationally bounded, and in the cases we chose it systematically overestimated the
binding energy; this contrasts from previous, similar work for nuclei, where RPA both
over- and underestimated the binding energy \cite{SJ02}. Because the overestimate
was not uniform, HF and PHF calculations sometimes gave accidentally better estimates
of the ionization potentials and electron affinities.
Perhaps most usefully, we found that for an example fixed atomic system, RPA tracks the CI binding
as we varied parameters for the basis, suggesting RPA might provide an efficient
proxy to speed up optimization of a basis. This will be explored further in future work.
In principle we can also compute and compare transitions and strength functions,
as has been done for nuclei \cite{SJ03}, and static and dynamic
observables such as static moments, polarizabilities, etc. This we also leave for future work.
We thank Michael Bromley for many valuable conversations and suggestions, and in particular for aid
in validating our codes.
The U.S.~Department of Energy supported this investigation through
grants DE-FG02-96ER40985 and DE-FC02-07ER41457.
|
\section{Introduction}
\label{sec-intro}
The picture of the nucleon as a system of interacting quarks and gluons naturally leads to the question about the intrinsic motion of these elementary particles inside the proton or neutron.
This intrinsic motion, specifically with respect to the transverse momentum,
can be described in terms of Transverse Momentum Dependent Parton Distribution Functions (TMDs\xspace), see, e.g.,
chapter 2 of Ref. \cite{Boer:2011fh} for a recent review.
TMDs for quarks, generically denoted by $f_1(x,\vprp{k}^2)$, $g_1(x,\vprp{k}^2)$, etc., encode essential information about the distribution of partons with respect to the longitudinal momentum fraction, $x$, and intrinsic quark transverse momentum, $\vprp{k}$.
With certain restrictions in mind, they have an intuitively appealing interpretation as three-dimensional probability densities \cite{Bacchetta:1999kz,Collins:2003fm}.
TMDs can, for example, be studied on the basis of angular asymmetries observed in processes such as Semi Inclusive Deep Inelastic Scattering (SIDIS) using suitable QCD factorization theorems that go beyond the standard collinear factorization,
see, e.g., Refs.~\cite{Ji:2004wu,Ji:2004xq,Aybat:2011zv,CollinsBook2011}.
In contrast to the usual collinear PDFs, TMDs turn out to be in general non-universal, i.e., process-dependent.
The process dependence arises from the difference in the final and initial state interactions in SIDIS and Drell-Yan scattering, respectively.
On the theoretical level, it can be understood as an intriguing consequence of the local color gauge invariance of the strong interaction
and the corresponding non-trivial gauge-link structures.
Specifically, QCD factorization leads to the remarkable prediction that the naively time reversal odd (T-odd) TMDs, in particular
the Sivers and Boer-Mulders functions, differ in sign for DY compared to SIDIS, $f^{\text{T-odd,SIDIS}}=-f^{\text{T-odd,DY}}$.
The implications and consequences of these observations continue to stir intense interest of many theoreticians and experimentalists,
as a number of fundamental questions and interesting puzzles remain to be addressed.
Motivated by promising experimental results from COMPASS, HERMES and JLab
(see, e.g., \cite{COMPASS:2008dn,HERMES:2009ti,Avakian:2010ae} and references therein),
as well as considerable progress on the theoretical and phenomenological sides during recent years,
an essential part of the physics program of future facilities will therefore be targeted in this direction,
including JLab $12\units{GeV}$ and the proposed EIC at JLab or BNL.
Theoretical calculations of TMDs\xspace from first principles require non-perturbative methods such as lattice QCD.
In previous works, we have introduced and explored techniques that allow the computation of the underlying amplitudes
on the lattice using non-local operators \cite{MuschThesis2,Hagler:2009mb,Musch:2010ka}.
Our numerical studies for a ``process-independent", direct gauge link geometry already produced encouraging results.
In this work, we present a first exploratory lattice study employing a more complex, ``process-dependent" link geometry
that gives us rather direct access to highly interesting T-odd observables.
In section \ref{sec-formalism}, we present the formalism and techniques
required for our calculations,
and provide definitions of the relevant T-odd and T-even TMD observables.
After a short introduction to the lattice computations at the beginning of section \ref{sec-latticecalc}, we continue with a presentation
and discussion of our numerical results for the generalized shifts and tensor charge.
A summary and conclusions are given in section \ref{sec-summary}.
\section{Formalism}
\label{sec-formalism}
\subsection{Definition of TMDs\xspace}
\label{sec-TMDdef}
In a relativistic quantum field theory, the question ``What is the probability to find a quark with a given momentum $k$ inside the proton?'' needs to be stated more precisely.
First of all, it turns out to be advantageous to formulate everything in light cone coordinates, see appendix \ref{sec-conv},
and to consider a frame of reference where the nucleon has large momentum in $z$-direction, i.e., $P^+ \gg m_N$, $\vprp{P} = 0$.
In light cone coordinates, the components $k^+$,~$\vprp{k}$,~$k^-$ of the quark momentum $k$ scale as $P^+/m_N$,~$1$,~$m_N/P^+$, respectively, under boosts along the $z$-axis. Thus the longitudinal momentum fraction of the quark $x = k^+ / P^+$ and its transverse momentum $\vprp{k}$ are invariant under boosts along the $z$-axis, while the $k^-$ component is suppressed.
This leads to the concept of transverse momentum dependent parton distribution functions (TMDs\xspace), which are functions of the longitudinal momentum fraction $x \equiv k^+ / P^+$ and of the quark transverse momentum $\vprp{k}$.
The transverse momentum components $\vprp{k}$ are particularly interesting, because they describe an intrinsic motion of the quarks inside the proton that occurs independent of the momentum of the proton itself. This gives us a unique picture of the dynamics inside the proton. Moreover, the TMDs\xspace are an important ingredient in our understanding of the origin of large angular- and spin-asymmetries found in experiments studying, e.g., semi-inclusive deep inelastic scattering (SIDIS) or the Drell-Yan process (DY).
In a naive approach based on a theory quantized on the light front, one obtains a momentum dependent number density of quarks from $f_1(x,\vprp{k}) \sim \frac{1}{2} \sum_{\Lambda = \pm 1} \sum_{\lambda = \pm 1} |a_{\lambda,q}(x,\vprp{k}) \ket{P,S}|^2$ (up to normalization factors), where $a_{\lambda,q}$ is an annihilation operator of quarks of flavor $q$ and helicity $\lambda$. The average over nucleon helicities $\frac{1}{2}\sum_{\Lambda \pm 1}$ implements an average over the spin $S$ in the nucleon state $\ket{P,S}$.
In this example, the TMD $f_1(x,\vprp{k})$ describes the distribution of unpolarized quarks in an unpolarized nucleon.
Rewriting the annihilation operator in terms of local quark field operators
$\bar q$ and $q$ reveals a problem: $f_1(x,\vprp{k})$ is a Fourier transform
of the matrix element $\bra{P,S} \bar q(0) \gamma^+ q(b) \ket{P,S}$ with
respect to the position $b$, and the bi-local operator
$\bar q(0) \gamma^+ q(b)$ is not gauge invariant, see
Ref. \cite{Collins:2003fm} for a review of the issue. Gauge invariance
can be restored by inserting a Wilson line $\WlineC{\mathcal{C}_b}$
between the quark fields, as defined in appendix \ref{sec-conv}. The Wilson
line introduces divergences that cannot be treated by conventional dimensional
regularization \cite{Collins:2008ht}. Several different schemes have been
proposed in the literature as to how to subtract those divergences
\cite{Collins:2003fm,Collins:2004nx,Ji:2004wu,Hautmann:2007uw,Chay:2007ty,Cherednikov:2008ua,Collins:2008ht,Aybat:2011zv,CollinsBook2011,Aybat:2011ge},
see Ref. \cite{Cherednikov:2011ft} for a recent comparison. In general,
these schemes require the introduction of a so-called soft factor
$\tilde \mathcal{S}$ inside the defining correlator of TMDs\xspace. The starting point
for our discussion of TMDs\xspace is thus a correlator of the general form
\begin{align}
\Phi^{[\Gamma]}(k,P,S;\ldots)\ \equiv\ \
\int \frac{d^4 b}{(2\pi)^4}\, e^{i k\cdot b}\,
\frac{\overbrace{\rule{0em}{1.2em}
\frac{1}{2}\, \bra{P,S}\ \bar{q}(0)\, \Gamma\ \WlineC{\mathcal{C}_b}\ q(b)\ \ket{P,S}
}^{\displaystyle \equiv \widetilde \Phi_\text{unsubtr.}^{[\Gamma]}(b,P,S;\ldots)}}
{ \widetilde{\mathcal{S}}(b^2;\ldots) }
\label{eq-corr}
\end{align}
The detailed properties of the Wilson line $\WlineC{\mathcal{C}_b}$ and the soft factor need to be specified by additional parameters, which we indicate by the dots ``$\ldots$'' for now and which will be discussed later. Moreover,
all objects above implicitly depend on a UV renormalization scale $\mu$.
In Eq.~\eqref{eq-corr}, $\tilde \mathcal{S}$ stands somewhat symbolically for an
expression that can, depending on the formalism, involve several vacuum
expectation values.
For example, in the scheme developed in Refs. \cite{Aybat:2011zv,CollinsBook2011,Aybat:2011ge}, our factor $\widetilde{\mathcal{S}}(b^2;\ldots)$ would be (using the notation of those references)
\begin{equation}
\widetilde{\mathcal{S}}(b^2;\ldots) =
\sqrt{ \frac
{ \tilde S_{(0)}( \vprp{b}, + \infty, -\infty )\ \tilde S_{(0)}( \vprp{b}, y_s, -\infty ) }
{ \tilde S_{(0)}( \vprp{b}, + \infty, y_s ) }
}
\end{equation}
where each of the objects $\tilde S_{(0)}( \vprp{b}, \ldots )$ is a
vacuum expectation value of Wilson line structures.
In this specific framework, the starting point of the discussion is space-like Wilson lines. Some Wilson lines remain tilted away from the light cone, leading to the dependence on the rapidity parameter $y_s$ in the above expression. Other Wilson lines, including those contained in $\WlineC{\mathcal{C}_b}$ in the numerator of Eq.~\eqref{eq-corr}, are brought back to the light cone in the sense of a limit, as indicated symbolically by $+\infty$ and $-\infty$ in the equation above, see Refs. \cite{Aybat:2011zv,CollinsBook2011} for details. As we will see in section \ref{sec-linkgeom} and below, certain matrix elements with space-like structures of Wilson lines are directly accessible on the Euclidean lattice, whereas taking the light-cone limit is only possible in the form of a numeric limit and technically challenging.
For the purposes of our treatment, however, we do not need to go into any
detail concerning the definition of $\widetilde{\mathcal{S}}(b^2;\ldots)$
in any particular framework,
since it will cancel in the observables we consider.
Integrating the correlator over the suppressed momentum
component $k^-$ yields
\begin{align}
\Phi^{[\Gamma]}(x,\vprp{k};P,S;\ldots) & \equiv \int dk^{-} \Phi^{[\Gamma]}(k,P,S;\ldots) \nonumber \\
& =\int \frac{d^2 \vprp{b}}{(2\pi)^2} \int \frac{d (b {\cdot} P)}{(2\pi) P^+}\ e^{i x (b {\cdot} P) - i\vprp{b} {\cdot} \vprp{k}}\,
\left. \frac{\frac{1}{2}\, \bra{P,S}\ \bar{q}(0)\, \Gamma\ \WlineC{\mathcal{C}_b}\ q(b)\ \ket{P,S}}
{ \widetilde{\mathcal{S}}(-\vprp{b}^2;\ldots) } \right|_{b^+ = 0} \ .
\end{align}
Notice that integrating over $k^-$ corresponds to setting $b^+=0$. As a consequence, $x \leftrightarrow (b {\cdot} P)$ and $\vprp{k} \leftrightarrow \vprp{b}$ act as independent pairs of Fourier conjugate variables in the expression above. The above correlator can be decomposed into TMDs\xspace.
For choices of the Dirac matrix $\Gamma$ that project onto leading twist, one obtains \cite{Ralston:1979ys,Tangerman:1994eh,Mulders:1995dh,Goeke:2005hb}
\begin{align}
\Phi^{[\gamma^+]}(x,\vprp{k};P,S,\ldots) & = f_1 - \toddmark{\frac{\ensuremath{\epsilon}_{ij}\, \vect{k}_{i}\, \vect{S}_{j}}{m_N}\ f_{1T}^\perp} \label{eq-phigammaplus}\ , \displaybreak[0] \\
\Phi^{[\gamma^+\gamma^5]}(x,\vprp{k};P,S,\ldots) & = \Lambda\, g_{1} + \frac{\vprp{k} \cdot \vprp{S}}{m_N}\ g_{1T} \label{eq-phigammaplusgfive} \ , \displaybreak[0] \\
\Phi^{[i\sigma^{i+}\gamma^5]}(x,\vprp{k};P,S,\ldots) & = \vect{S}_i\ h_{1} + \frac{(2 \vect{k}_i \vect{k}_j - \vprp{k}^2 \delta_{ij}) \vect{S}_j}{2 m_N^2}\, h_{1T}^\perp + \frac{\Lambda \vect{k}_i}{m_N} h_{1L}^\perp + \toddmark{\frac{\ensuremath{\epsilon}_{ij} \vect{k}_{j}}{m_N} h_1^\perp}\ .\label{eq-phisigmaplusi}
\end{align}
The TMDs\xspace $f_1$, $g_1$, $h_1$, $g_{1T}$, $h_{1L}^\perp$, $h_{1T}^\perp$, $f_{1T}^\perp$ and $h_1^\perp$ are functions of $x$, $\vprp{k}^2$, $\mu$ and further parameters related to regularization and link geometry.
The structures shown in brackets $[\ ]_\text{odd}$ involve so-called naively time reversal odd (T-odd) TMDs\xspace, namely the Sivers function $f_{1T}^\perp$ \cite{Sivers:1989cc} and the Boer-Mulders function $h_1^\perp$ \cite{Boer:1997nt}. The origin of the above parametrization and the special role of T-odd TMDs\xspace will become clear after we have discussed the geometry of the gauge link path
$\mathcal{C}_b $ and symmetry transformation properties.
\subsection{General strategy}
At this point, several remarks are in order as to how we aim to introduce TMD\xspace observables that can be accessed with lattice QCD using a non-local operator technique.
Due to the underlying operator structure, the situation is quite different from that of standard collinear
PDFs and offers unique opportunities and challenges.
As an introductory example, consider the definition of a standard PDF in the unpolarized case,
\begin{equation*}
f_1(x) \equiv \frac{1}{2 (2\pi)} \int db^- e^{i x P^+ b^-} \, \bra{P,S}\ \bar{q}(0)\, \gamma^+\ \Wline{0,n b^-}\ q(n b^-)\ \ket{P,S} \ .
\end{equation*}
For PDFs, the gauge link $\Wline{0,n b^-}$ is simply a straight, light-like Wilson line of finite extent connecting the two quark field operators \cite{Collins:1981uw}. No continuous Lorentz transformation exists that allows us to ``rotate'' the non-local operator $\bar{q}(0)\, \gamma^+\ \Wline{0,n b^-}\ q(n b^-)$ into Euclidean space. The light-like separation stays always light-like, but in Euclidean space objects cannot have any extent in (Minkowski-) time.
As a consequence, one is forced
to invoke the operator product expansion to cast the calculation in
terms of local matrix elements which can be accessed using lattice QCD.
The situation for TMDs\xspace differs fundamentally in several aspects:
\begin{enumerate}
\item The separation $b$ of the quark field operators has an additional transverse component, $b = n b^- + b_\perp$. Thus, in general, this separation is space-like. This opens the possibility of a direct representation of the non-local operator in Euclidean space.
\item The geometry of the gauge link $\WlineC{\mathcal{C}_b}$ is more complicated, depends to a certain degree on the experiment under consideration and in general extends out to infinity. As a result, it becomes questionable whether an expansion in terms of local operators is possible at all.
\item Regularization is more complicated, leading to the introduction of the soft factor $\tilde \mathcal{S}$ and additional regularization parameters beyond the usual renormalization scale $\mu$ of the $\overline{\mathrm{MS}}$ scheme.
\end{enumerate}
The first two items listed above are our main motivation to develop a technique for lattice studies of TMDs\xspace based on non-local operators.
It should be emphasized that this technique can only work for the analysis of certain TMD\xspace-related observables within a limited kinematical range. The method cannot be applied to study the $x$-dependence of PDFs directly, without the use of non-trivial extrapolations.
In previous publications \cite{Hagler:2009mb,Musch:2010ka}, it was
demonstrated that the non-local operator technique is quite promising and
produces interesting results, for a simplified gauge link geometry, at
least on a qualitative level. The crucial connection between the formalism
in Minkowski space and the results from Euclidean space is provided through
a parametrization in terms of invariant amplitudes\footnote{Note that the
symbol $l$ in Ref. \cite{Hagler:2009mb,Musch:2010ka} corresponds to
$-b$ in the present study.} $\widetilde{A}_i(b^2,b {\cdot} P)$. By
virtue of their Lorentz-invariance, the calculation of these amplitudes
can be performed in any desired Lorentz frame. In particular, for the
generic off-light cone kinematics appropriate for TMDs\xspace, there is no
obstacle to performing the calculation in a frame in which the nonlocal
operator in question is defined entirely at one fixed time. In this frame,
one can cast the computation of the nonlocal matrix element
in terms of a Euclidean path integral, evaluated employing the standard
methods of lattice QCD.
The study at hand builds directly on Ref. \cite{Musch:2010ka}, and we refer the reader to that publication for an introduction to the essential principles of the methology. One of the remaining challenges identified in Ref. \cite{Musch:2010ka} concerns the geometry of the gauge link. In the present study, we replace the simple straight connection by a staple-like path that corresponds more accurately to the situation in phenomenology. We stress that these gauge link
structures are part of the established phenomenological framework, which
we take as given, and not a new assumption related to our use of
lattice QCD as a calculational method. Whereas our results depend on
the gauge link structure, specific physical processes such as SIDIS and
DY unambiguously correspond to definite instances of that structure.
Throughout our discussion, we clearly identify the SIDIS and DY limits
of our data.
It is important to point out that our assumptions about the operator
structure of TMDs\xspace rely on factorization arguments that are much more
involved than for the usual PDFs. In fact, one must be judicious
concerning the classes of reactions for which it can be assumed that
a factorization framework with well-defined TMDs\xspace exists. For example,
it has been realized recently \cite{Rogers:2010dm,Aybat:2011vb}
that TMD factorization generally fails for large reaction classes,
in particular processes with multiple hadrons in both the initial and
the final state. While the consequences of this observation are not yet
all known, it appears certain that to develop a TMD framework for these
processes, a fundamental change of perturbative QCD techniques is needed.
It could, e.g., very well be that measurement-independent cross-sections
are simply not defined for certain reaction classes and that, instead,
the appropriate quantities will be entanglement amplitudes which then
have to be folded with quantities encoding the measurement
process \cite{RogersPriv}. In contradistinction to the aforementioned
classes of reactions, for other types of processes such as SIDIS and DY,
recent progress \cite{Aybat:2011zv,CollinsBook2011,Aybat:2011ge} indicates
that a valid definition of TMDs\xspace based on factorization arguments indeed is
possible, within a scheme regularized employing space-like links. Promising
steps have been taken to develop the predictive capabilities of this
framework \cite{Aybat:2011ta}. A pertinent discussion is given in
section I of our previous publication \cite{Musch:2010ka}, with further
details to be found in the references therein and recent overviews in
Refs.~\cite{Cherednikov:2011ft,Aybat:2011vb}.
The point which we wish to emphasize here is that it is not the purpose
of our present work to critique or justify the various approaches to
defining TMDs\xspace in terms of operators and matrix elements which have been
advanced in response to issues of factorization and regularization.
Instead, we will assume
that a good definition of TMDs\xspace with a connection to phenomenology through
a valid factorization argument exists for certain classes of processes
such as SIDIS and DY; we focus exclusively on those TMDs\xspace and do not aim
to contribute to discussions of factorization, fragmentation functions,
or related matters. Our starting point thus is the definition of TMDs\xspace in
terms of a TMD\xspace correlator of the rather general form \eqref{eq-corr}.
Working from this definition, we will moreover restrict ourselves to
observables in which the soft factor cancels, so that specifics of the
soft factor are not relevant for our results. Regarding the
necessary regularization of the gauge link, we pick the proposal that is
most suitable for our purposes, namely tilting the gauge link slightly
away from the light cone \cite{Collins:1981uk}, in a space-like direction
\cite{Aybat:2011zv,CollinsBook2011}. We stress that, in choosing this
approach to defining and regularizing TMDs\xspace, we are led to consider kinematics
off the light cone from the very beginning, which makes a connection to
Euclidean lattice QCD feasible, as already noted further above. We emphasize
that, within this work, we do not aim to arrive at any statements concerning
the formal nature of the light-cone limit. We will, however, focus particularly
on the behavior of our numerical results as we extend the kinematic
region as far towards the light cone as possible.
One necessary step of a lattice calculation is to discretize the operators.
The discretization of non-local operators as we encounter them here is still
a rather new concept. An important assumption we make is that non-local
lattice operators composed of structures much larger than the lattice
spacing essentially renormalize in the same fashion as their counterparts
in the continuum, except that the renormalization parameters are specific
to the lattice action and the discretization prescription. We have given
reasons for this assumption and explored it numerically in
Ref.~\cite{Musch:2010ka}, see in particular sections III~D, IV~B, IV~C and
appendices B, D, G and H therein. However, we point out that a more rigorous
treatment would still be desirable. Especially the question of mixing
properties as one attempts to make contact with the local operator
formalism remains a challenge for the future.
Keeping the above remarks in mind, it is worthwhile summarizing the logic
underlying our treatment succinctly before laying out the details further
below:
\begin{enumerate}
\item
We start from a definition of TMDs\xspace in terms of the correlator \eqref{eq-corr},
considering generic off-light cone kinematics from the very beginning.
\item
The correlator \eqref{eq-corr} is parametrized in terms of Lorentz-invariant
amplitudes, cf.~section~\ref{sec-parametr} below. This crucial step permits
one to transform results into different Lorentz frames in a simple manner.
\item
On this basis, we choose the Lorentz frame in which the nonlocal operator
entering \eqref{eq-corr} is defined at one single time as the one most
suitable for our calculation. We stress again that there is no obstacle
to this choice, since the separations in the operator are all space-like.
\item
In the aforementioned frame, the computation of the nonlocal matrix element
can be cast in terms of a Euclidean path integral and performed employing
the standard methods of lattice QCD.
\item
We form appropriate ratios of the extracted invariant amplitudes in which
soft factors and multiplicative renormalization factors cancel, such as
the $\vprp{k}$-shifts discussed in section~\ref{sec-latquan}, which, in
principle, represent measurable quantities. We particularly study the
approach to the SIDIS and DY limits in these quantities.
\end{enumerate}
\subsection{TMDs\xspace in Fourier space and $x$-integration}
In essence, the lattice method we use allows us to evaluate the $b$-dependent matrix elements $\widetilde \Phi_\text{unsubtr.}^{[\Gamma]}(b,P,S;\ldots)$ introduced in Eq. \eqref{eq-corr}. As a result, it is more direct and natural to state our results in terms of Fourier-transformed, $\vprp{b}$-dependent TMDs\xspace and their $\vprp{b}$-derivatives. For a generic TMD\xspace $f$ we define
\begin{align}
\tilde f(x, \vprp{b}^2;\ldots) & \equiv \int d^2 \vprp{k} \, e^{i \vprp{b} \cdot \vprp{k} }\; f(x, \vprp{k}^2;\ldots)
= 2\pi \int d|\vprp{k}| |\vprp{k}|\ J_0(|\vprp{b}||\vprp{k}|)\ f(x, {\vprp{k}^2};\ldots )\ , \label{eq-FTMD} \\
\tilde f^{(n)}(x, \vprp{b}^2\ldots ) & \equiv n!\left( -\frac{2}{m_N^2}\partial_{\vprp{b}^2} \right)^n \
\tilde f(x, \vprp{b}^2;\ldots )
= \frac{ 2\pi \ n!}{(m_N^2)^n} \int d |\vprp{k}| |\vprp{k}|\left( \frac{|\vprp{k}|}{|\vprp{b}|}\right)^n J_n(|\vprp{b}||\vprp{k}|)\ f(x, {\vprp{k}^2};\ldots )\ , \label{eq-dFTMD}
\end{align}
where the $J_n$ are Bessel functions of the first kind, and $m_N$ is the mass of the target hadron.
These objects and their potential phenomenological relevance have been discussed in detail in Ref. \cite{Boer:2011xd}. Moreover, evolution equations are naturally expressed in terms of the $\tilde f^{(n)}$, compare, e.g., Ref.~\cite{Idilbi:2004vb}. In the limit $|\vprp{b}| \rightarrow 0$, one recovers conventional $\vprp{k}$-moments of TMDs\xspace:
\begin{align}
\tilde f^{(n)}(x, 0;\ldots ) & = \int d^2 \vprp{k} \left( \frac{\vprp{k}^2}{2 m_N^2 }\right)^n f(x, \vprp{k}^2;\ldots ) \equiv f^{(n)}(x)\ .
\label{eq-btder2}
\end{align}
However, it is known \cite{Bacchetta:2008xw} that $\vprp{k}$-moments like $f_1^{(0)}(x)$ and $f_{1T}^{\perp(1)}(x)$ are ill-defined without further regularization. The problem is that the integral in the above equation diverges if the integrand does not fall off quickly enough in the region of large $\vprp{k}$, where the TMDs\xspace $f(x,\vprp{k}^2)$ are perturbatively predictable. Even though $\vprp{k}$-moments may be more familiar to the reader, we therefore do not attempt to extrapolate to $\vprp{b}=0$, but rather state our results at finite $|\vprp{b}|$, where the $\vprp{k}$-integrals of Eqs. \eqref{eq-FTMD} and \eqref{eq-dFTMD} can be shown to be convergent in the relevant cases \cite{Boer:2011xd}.
Information about the $x$-dependence of TMDs\xspace can be obtained from the lattice via the Fourier-conjugate variable, $b {\cdot} P$ \cite{Hagler:2009mb,Musch:2010ka}.
However, the calculations performed in Euclidean space only allow us to access a limited range of $b {\cdot} P$, precluding us from performing a straightforward Fourier transform.
In this work, we limit ourselves to the study of $x$-integrated TMDs\xspace
\begin{align}
f^{[1]}(\vprp{k}^2;\ldots) \equiv \int_{-1}^1 dx\ f(x,\vprp{k}^2;\ldots)\ .
\end{align}
These are accessible from the data at $b {\cdot} P=0$.
Here, the superscript ${}^{[1]}$ denotes the first Mellin moment in $x$.
The integration is performed over the full range of $x$.
TMDs\xspace evaluated at negative values of $x$ can be related to anti-quark distributions, see, e.g., \cite{Mulders:1995dh,Musch:2010ka} for details.
\subsection{Quantities suitable for lattice extraction}
\label{sec-latquan}
Certain ratios of $\vprp{k}$-moments of TMDs\xspace have interesting physical interpretations. For example, consider
\begin{align}
m_N \frac{f_{1T}^{\perp(1)}(x)}{f_1^{(0)}(x)} =
\left. \frac{ \int d^2 \vprp{k}\, \vect{k}_y \ \Phi^{[\gamma^+]}(x,\vprp{k},P,S;\ldots) }{ \int d^2 \vprp{k} \phantom{\vect{k}_y}\ \Phi^{[\gamma^+]}(x,\vprp{k},P,S;\ldots) } \right|_{\displaystyle \vprp{S}=(1,0) }\, ,
\end{align}
where $\gamma^+$ projects on leading-twist.
In the context of the density interpretation of TMDs\xspace mentioned in section \ref{sec-TMDdef}, the ratio above yields the average transverse momentum
in $y$-direction, for quarks with given longitudinal momentum fraction $x$ inside a proton polarized in $x$-direction.
We will show below that quantities like this can be calculated rather directly on the lattice.
For the reasons mentioned above, we limit ourselves to ratios formed from $x$-integrated quantities. Let us therefore consider
\begin{align}
\langle \vect{k}_y \rangle_{TU} \equiv m_N \frac{f_{1T}^{\perp[1](1)}}{f_1^{[1](0)}} \ .
\label{eq-sivshift}
\end{align}
Ignoring the role of anti-quarks, this ratio, called in the following ``Sivers shift'', represents the average transverse momentum of unpolarized (``U'') quarks orthogonal to the transverse (``T'') spin of the nucleon.
Note, however, that the denominator $f_1^{[1](0)}$ arises from a difference of quarks and anti-quarks and thus
gives the number of valence quarks in the nucleon.
On the other hand, in the numerator $f_{1T}^{\perp[1](1)}$, the average transverse momentum of quarks and anti-quarks is summed over \cite{Mulders:1995dh,Musch:2010ka}.
A profound interpretation of $f_{1T}^{\perp[1](1)}$ in impact parameter space has been given in Ref.~\cite{Burkardt:2003uw}.
However, as mentioned before, understanding $f_{1T}^{\perp[1](1)}$ simply as a $\vprp{k}$-weighted TMD\xspace is problematic, since the $\vprp{k}$-integral is expected to be UV divergent.
A natural way of circumventing this divergence is to generalize the Sivers shift to an expression in terms of the Fourier-transformed TMDs\xspace:
\begin{align}
\langle \vect{k}_y \rangle_{TU}(\vprp{b}^2;\ldots) \equiv m_N \frac{\tilde f_{1T}^{\perp[1](1)}(\vprp{b}^2;\ldots)}{\tilde f_1^{[1](0)}(\vprp{b}^2;\ldots)} \ .
\end{align}
This is the type of quantity that we investigate in the present study. In the limit $\vprp{b}^2=0$ we recover the Sivers shift \eqref{eq-sivshift}, because the Fourier transformed TMDs\xspace $\tilde f_{1T}^{\perp[1](1)}$ and $\tilde f_1^{[1](0)}$ coincide with the moments $f_{1T}^{\perp[1](1)}$ and $f_1^{[1](0)}$, respectively. We are, however, interested in the generalized Sivers shift for non-zero $\vprp{b}^2$, where the said UV-divergence disappears. The variable $\vprp{b}^2$ effectively acts as a regulator. Moreover, the $\vprp{b}$-dependence allows us to study differences in the widths of distributions on a qualitative level.
\subsection{Link geometry}
\label{sec-linkgeom}
The prescription for the geometry of the gauge link path $\mathcal{C}_b$ affects both the number of allowed structures appearing in Eqs. \eqref{eq-phigammaplus}-\eqref{eq-phisigmaplusi} and the numerical result for the TMDs\xspace. We therefore need to ask which link geometries are appropriate.
The simplest link geometry is a straight line connecting the quark fields at $0$ and $b$, see Fig. \ref{fig-link-straight}. TMDs\xspace with straight gauge links have been studied on the lattice in Refs. \cite{Hagler:2009mb,Musch:2010ka}. While these ``process-independent'' TMDs\xspace are interesting from a theoretical point of view in their own right, it is so far not known how to relate these quantitatively to the TMDs\xspace that play a role in scattering experiments. The operator with straight gauge links offers the largest possible degree of symmetry. As a result, T-odd TMDs\xspace vanish for straight gauge links.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-link-straight
\includegraphics[trim=0 0 120 0,clip=true]{figs/DirectLink}
}\qquad%
\subfloat[][]{%
\label{fig-link-staple}%
\includegraphics[]{figs/StapleLink2}
}%
\caption[SIDIS diagram]{%
\subref{fig-link-straight}\ %
Straight gauge link.\ \ \par
\subref{fig-link-staple}\ %
Staple-shaped gauge link as in SIDIS and DY.\par%
\label{fig-links}%
}
\end{figure}
For TMDs\xspace that allow us to describe measurable effects in scattering experiments such as SIDIS or DY, the form of the gauge link is largely dictated by the physical process. To understand scattering experiments at high momentum transfer $Q$, one tries to apply approximations valid for large $Q$ that separate hard, perturbative and soft, non-perturbative scales in the dominant physical processes in order to arrive at an expression for the cross section in factorized form.
In the standard collinear approximation, all internal transverse momenta are integrated out and conventional parton distribution functions and fragmentation functions are used to describe the process. In certain kinematical regions this approximation is insufficient. An example is SIDIS, where the momentum $P_h$ of one of the final state hadrons is measured after a lepton-nucleon collision at large momentum transfer $Q$. The transverse momentum dependent formalism is needed when the transverse momentum component $\vect{P}_{h\perp}$ is small with respect to $Q$, see, e.g., Ref. \cite{Bacchetta:2008xw} for an in-depth discussion.
The leading diagram for SIDIS
is shown in a simplified, factorized form in Fig. \ref{fig-SIDISdiagram}.
The lower shaded bubble in the diagram represents the structure parametrized by TMDs\xspace. A gauge link in the TMD\xspace correlator arises naturally as an idealized, effective, resummed description of the gluon exchanges between the ejected quark and the remainder of the nucleon in the evolving final state, see, e.g., Ref.~\cite{Pijlman:2006tq} for a review.
\begin{figure}[btp]
\centering%
\includegraphics[width=0.5\linewidth]{figs/SIDISdiagram4_full.pdf}
\caption{%
Illustration of the leading contribution to SIDIS in factorized form.
\label{fig-SIDISdiagram}
}
\end{figure}
The gauge link roughly follows the direction of the ejected quark, in SIDIS by convention denoted by the light-cone $\ensuremath{n}$ direction. The TMD\xspace correlator obtained from the squared amplitude thus has parallel Wilson lines attached to each of the quark field operators at $0$ and $b$, extending out to infinity along a direction $v \approx \ensuremath{n}$, see Fig. \ref{fig-link-staple}. Due to the fact that the gauge link is only an effective representation of final state interactions within a framework of suitable approximations, there is a certain degree of freedom with respect to its geometry, in particular with regard to the choice of its direction $v$. At tree level, the most convenient choice is an exactly light-like gauge link, $v = \ensuremath{n}$. However, going beyond tree-level, it has been found that the light-like link introduces so-called rapidity divergences that are hard to remove, see Ref. \cite{Collins:2007ph} for a review. One way of regulating these divergences is to use a gauge link slightly off the light cone \cite{Collins:1981uw}, see Refs. \cite{Ji:2004wu,Aybat:2011zv,CollinsBook2011} for the application to SIDIS. In Ref. \cite{Ji:2004wu}, the direction $v$ is chosen time-like. More recent work in Refs. \cite{Aybat:2011zv,CollinsBook2011} is based on space-like Wilson lines, motivated by the insight that TMDs\xspace with this choice of link directions feature a ``modified universality'', i.e., they are predicted to be numerically equal for both SIDIS and DY \cite{Collins:2004nx} up to the expected sign changes of T-odd TMDs\xspace. The space-like choice of Wilson lines also opens up the possibility of implementing the gauge link directly in lattice QCD.
In Fig. \ref{fig-link-staple}, the two parallel Wilson lines are connected at the far end by another straight Wilson line. The complete gauge link thus has a staple-like shape. Bridging the transverse gap is necessary to render the operator gauge invariant and proves to be essential if the light-cone gauge $\ensuremath{n} \cdot A = 0$ is used \cite{Belitsky:2002sm,Brodsky:2002cx}. In a covariant gauge, the connecting link at infinity can be omitted; this has been exploited in Refs.~\cite{Ji:2004wu,Aybat:2011zv}. Lattice calculations are typically performed without any gauge fixing. We therefore prefer the notation with an explicitly gauge invariant operator. Moreover, in our study we take the limit of an infinite ``staple extent'' $\eta$ explicitly. The gauge link employed in this work thus reads
\begin{align}
\WlineC{\mathcal{C}_b^{(\eta v)}} = \Wline{0, \eta v, \eta v + b, b}\ ,
\label{eq-staplelink}
\end{align}
where $v$ is space-like. Even at finite $\eta$, this gauge link geometry fulfills the desired symmetry transformation rules, as listed in Eq. (C6) of Ref. \cite{Musch:2010ka} and discussed further below.
Here, we will be mostly concerned with the lowest $x$-moment of TMDs\xspace, corresponding to the case $b^- = b^+ = 0$.
In this case, the connection at the far end is purely transverse.
We choose $v$ space-like and, as in Refs.~\cite{Ji:2004wu,Aybat:2011zv,CollinsBook2011}, we consider TMDs\xspace for the choice $\vprp{v}=0$.
The Lorentz-invariant quantity characterizing the direction of $v$ is the parameter $\zeta \equiv 2 v {\cdot} P/\sqrt{|v^2|}$.
The light-like direction $v = \ensuremath{n}$ can be approached in the limit $\zeta \rightarrow \infty$.
The parameter $\zeta$ can be understood as an artificial scale or cutoff introduced to regulate rapidity divergences.
Within their work on $e^+e^-$-scattering, Collins and Soper provided evolution equations for the dependence on $\zeta$ applicable for $\zeta \gg \Lambda_{QCD}$ \cite{Collins:1981uk}.
Similar equations have been worked out for all leading-twist and spin-dependent parton distributions \cite{Idilbi:2004vb} based on the formalism of Ref. \cite{Ji:2004wu}.
For the more recent formalism of Refs. \cite{Aybat:2011zv,CollinsBook2011}, evolution equations are presently available for the unpolarized case and the Sivers function \cite{Aybat:2011ge}.
The vectors $P$ and $v$ can be written in terms of rapidities $y_P$ and $y_v$, respectively:
$P^{\pm} = m_N e^{\pm y_P} / \sqrt{2}$ and $v^+/v^- = - e^{2y_v}$.
Rewriting $\zeta$ as a dimensionless quantity,
\begin{equation}
\hat \zeta \equiv \zeta / 2 m_N = \frac{v \cdot P}{\sqrt{|v^2|} \sqrt{P^2}} = \sinh(y_P - y_v),
\label{eq-zeta}
\end{equation}
reveals that it is essentially a rapidity difference.
Notice that the entire system can always be boosted to a frame where $v$ has only spatial components, $v^0 = 0, y_v=0$. This is crucial for the lattice approach.
\subsection{Parametrization of the correlator}
\label{sec-parametr}
The translation of our results obtained in Euclidean space into TMDs\xspace defined and interpreted in the context of light cone coordinates is mediated through a parametrization of the correlator $\tilde \Phi_\text{unsubtr.}$ in terms of manifestly Lorentz-invariant amplitudes.
For our purposes, it will be important to take the dependence on the link direction $v$ explicitly into account.
A parametrization of the correlator
$\Phi_\text{unsubtr.}(k,P,S;\infty v,\mu) = \int d^4 b/(2\pi)^4\, e^{i k\cdot b}\, \frac{1}{2}\, \bra{P,S}\ \bar{q}(0)\, \Gamma\ \WlineC{\mathcal{C}_b^{(\infty v)}}\, q(b)\ \ket{P,S}$
for link paths that extend to infinity into a direction $v$
has been worked out in Ref. \cite{Goeke:2005hb} and involves 32 independent amplitudes $A_i$ and $B_i$ that depend on the Lorentz-invariant quantities
$k^2$, $k \cdot P$, $k {\cdot} v/v {\cdot} P$ and $\hat \zeta$.
Appendix C of Ref. \cite{Boer:2011xd} shows that a parametrization of the corresponding $b$-dependent correlator $\tilde \Phi_\text{unsubtr.}$ is of the same form as the parametrization of $\Phi_\text{unsubtr.}$ if we substitute $k \rightarrow - i m_N^2 b$. We thus obtain
\begin{align}
\frac{1}{2}\widetilde \Phi^{[\mathds{1}]}_{\text{unsubtr.}} & =
m_N \widetilde{A}_1
- \frac{i m_N^2}{v {\cdot} P} \epsilon^{\mu\nu\rho\sigma} P_\mu b_\nu v_\rho S_\sigma \widetilde{B}_5
\label{eq-phitildedecomp-scal} \displaybreak[0] \\
%
\frac{1}{2}\widetilde \Phi^{[\gamma^5]}_{\text{unsubtr.}} & =
m_N^2 (b {\cdot} S) \widetilde{A}_5
+ \frac{i m_N^2}{P {\cdot} v}(v {\cdot} S) \widetilde{B}_6
\label{eq-phitildedecomp-pscal} \displaybreak[0] \\
%
\frac{1}{2}\widetilde \Phi^{[\gamma^\mu]}_{\text{unsubtr.}} & =
P^\mu\, \widetilde{A}_2 - i m_N^2 b^\mu\, \widetilde{A}_3
- i m_N \epsilon^{\mu \nu \alpha \beta} P_\nu b_\alpha S_\beta\, \widetilde{A}_{12} \nonumber \\ &
+ \frac{m_N^2}{(v {\cdot} P)} v^\mu\, \widetilde{B}_1
+ \frac{m_N}{v {\cdot} P} \epsilon^{\mu \nu \alpha \beta} P_\nu v_\alpha S_\beta\, \widetilde{B}_7
- \frac{ i m_N^3}{v {\cdot} P} \epsilon^{\mu \nu \alpha \beta} b_\nu v_\alpha S_\beta\, \widetilde{B}_8 \nonumber \\ &
- \frac{m_N^3}{v {\cdot} P} (b {\cdot} S) \epsilon^{\mu \nu \alpha \beta} P_\nu b_\alpha v_\beta\, \widetilde{B}_9
- \frac{i m_N^3}{(v {\cdot} P)^2} (v {\cdot} S) \epsilon^{\mu \nu \alpha \beta} P_\nu b_\alpha v_\beta \widetilde{B}_{10}
\label{eq-phitildedecomp-vec} \displaybreak[0] \\
%
\frac{1}{2}\widetilde \Phi^{[\gamma^\mu \gamma^5]}_{\text{unsubtr.}} & =
- m_N S^\mu \widetilde{A}_6 + i m_N (b {\cdot} S) P^\mu \widetilde{A}_7 + m_N^3 (b {\cdot} S) b^\mu \widetilde{A}_8 \nonumber \\ &
+ \frac{i m_N^2}{v {\cdot} P} \epsilon^{\mu \nu \rho \sigma} P_\nu b_\rho v_\sigma \widetilde{B}_4
- \frac{m_N}{v {\cdot} P}(v {\cdot} S) P^\mu \widetilde{B}_{11}
+ \frac{i m_N^3}{v {\cdot} P}(v {\cdot} S) b^\mu \widetilde{B}_{12} \nonumber \\ &
+ \frac{i m_N^3}{v {\cdot} P}(b {\cdot} S) v^\mu \widetilde{B}_{13}
- \frac{m_N^3}{(v {\cdot} P)^2}(v {\cdot} S) v^\mu \widetilde{B}_{14}
\label{eq-phitildedecomp-pvec} \displaybreak[0] \\
%
\frac{1}{2}\widetilde \Phi^{[i \sigma^{\mu\nu} \gamma^5]}_{\text{unsubtr.}} & =
i m_N \epsilon^{\mu\nu\rho\sigma} P_\rho b_\sigma \widetilde{A}_4
+ P^{[\mu} S^{\nu]} \widetilde{A}_9
- i m_N^2 b^{[\mu} S^{\nu]} \widetilde{A}_{10}
- m_N^2 (b {\cdot} S) P^{[\mu} b^{\nu]} \widetilde{A}_{11} \nonumber \\ &
- \frac{ m_N}{v {\cdot} P} \epsilon^{\mu\nu\rho\sigma} P_\rho v_\sigma \widetilde{B}_2
+ \frac{im_N^3}{v {\cdot} P} \epsilon^{\mu\nu\rho\sigma} b_\rho v_\sigma \widetilde{B}_3
+ \frac{m_N^2}{v {\cdot} P} v^{[\mu} S^{\nu]} \widetilde{B}_{15}
- \frac{i m_N^2}{v {\cdot} P} (b {\cdot} S) P^{[\mu} v^{\nu]} \widetilde{B}_{16} \nonumber \\ &
- \frac{m_N^4}{v {\cdot} P} (b {\cdot} S) b^{[\mu} v^{\nu]} \widetilde{B}_{17}
- \frac{i m_N^2}{v {\cdot} P} (v {\cdot} S) P^{[\mu} b^{\nu]} \widetilde{B}_{18}
+ \frac{m_N^2}{v {\cdot} P} (v {\cdot} S) P^{[\mu} v^{\nu]} \widetilde{B}_{19}
- \frac{i m_N^4}{(v {\cdot} P)^2} (v {\cdot} S) b^{[\mu} v^{\nu]} \widetilde{B}_{20}
\label{eq-phitildedecomp-tens}
\ ,
\end{align}
where $a^{[\mu}b^{\nu]} \equiv a^\mu b^\nu - a^\nu b^\mu$. The structures above are compatible with the transformation properties of the correlator under the symmetries of QCD. For completeness we list them again in appendix \ref{sec-symtraf}.
Our previous studies of TMDs\xspace on the lattice \cite{Hagler:2009mb,Musch:2010ka} were carried out with straight gauge links. In that case only the T-even structures involving amplitudes of type $\widetilde{A}_i$ appear in the parametrization. As pointed out already in those references,
there is not necessarily a one-to-one correspondence between the $A_i$ and $\widetilde{A}_i$ (or the $B_i$ and $\widetilde{B}_i$).
For example, $\widetilde{A}_8$ contributes to $A_6$, $A_7$ and $A_8$. Note that $l=-b$ in Refs. \cite{Hagler:2009mb,Musch:2010ka}.
In the above parametrization, factors of $(v {\cdot} P)^{-n}$ ensure that the structures are invariant under rescaling of $v$, i.e., $v \rightarrow \alpha v$, for any $\alpha > 0$.
The above parametrization is therefore suitable for describing the case of the staple links extending to infinity. In that case, only the directional information contained in $v$ should enter. For the lattice calculations, it is however advantageous to start with an equivalent parametrization in which the structures explicitly depend on the staple extent $\eta$ and which is still well-defined for $v {\cdot} P = 0$. Such a parametrization can be obtained from the parametrization above by replacing $v \rightarrow \eta v$ and by leaving out the factors $(v {\cdot} P)^{-n}$. For example,
\begin{align}
\frac{1}{2}\widetilde \Phi^{[\mathds{1}]}_{\text{unsubtr.}}(b,P,S,\eta v,\mu) & =
m_N \tilde a_1
- i m_N^2 \epsilon^{\mu\nu\rho\sigma} P_\mu b_\nu \eta v_\rho S_\sigma \tilde b_5
\label{eq-phitildedecomp-scal-2} \ ,
\end{align}
and analogously for the other Dirac structures. Here we have used lower case amplitudes to distinguish the two parametrizations. The relation to the upper case amplitudes is given by
\begin{align}
\widetilde{A}_i\left(b^2,b {\cdot} P,\frac{v {\cdot} b}{v {\cdot} P}, \frac{v^2}{(v {\cdot} P)^2}, \eta v {\cdot} P\right) & = \tilde a_i(b^2,b {\cdot} P, \eta v {\cdot} b, (\eta v)^2, \eta v {\cdot} P) \, , \nonumber \\
\widetilde{B}_i\left(b^2,b {\cdot} P,\frac{v {\cdot} b}{v {\cdot} P}, \frac{v^2}{(v {\cdot} P)^2}, \eta v {\cdot} P\right) & = (\eta v {\cdot} P)^n\ \tilde b_i(b^2,b {\cdot} P, \eta v {\cdot} b, (\eta v)^2, \eta v {\cdot} P) \, ,
\label{eq-ab}
\end{align}
where $n$ is the power with which $v {\cdot} P$ appears in the denominator in front of the corresponding amplitude $\widetilde{B}_i$ in the parametrization.
Notice that the $\tilde a_i$ and $\tilde b_i$ are functions of all the Lorentz-invariant products of $b$, $P$ and $\eta v$.
For the upper case amplitudes, however, we choose to represent the dependence on these invariants in the third and fourth argument by $\eta$-independent expressions, in order to facilitate taking the limit $\eta \rightarrow \pm \infty$.
The dependence on the Collins-Soper parameter $\hat\zeta$ is given by the fourth argument, $v^2/(v {\cdot} P)^2 = - 1 /(m_N\hat\zeta)^2$,
while the fifth argument, $\eta v {\cdot} P$, characterizes the length of the gauge link and distinguishes between future and past pointing
Wilson lines.
For the calculation of TMDs\xspace we work in a frame with $b^+=0$ and $\vprp{v}=\vprp{P}=0$. This leads to a relation that can be expressed in Lorentz-invariant form as
\begin{align}
\frac{v {\cdot} b}{v {\cdot} P} = b {\cdot} P \frac{R({\hat \zeta}^2)}{m_N^2}\, ,
\label{eq-lindeprel}
\end{align}
where
\begin{align}
R({\hat \zeta}^2) \equiv 1- \sqrt{1 + \hat \zeta^{-2}} = \frac{m_N^2}{v {\cdot} P} \frac{v^+}{P^+} \, .
\end{align}
The relation Eq. \eqref{eq-lindeprel} shows that the third argument of the $\widetilde{A}_i$ and $\widetilde{B}_i$ is not independent of the others in the context of TMDs\xspace.
Moreover, in our lattice calculations, we have to choose the link directions $b$ and $v$ such that Eq. \eqref{eq-lindeprel} is fulfilled.
As a side remark, the parameter corresponding to Eq. \eqref{eq-lindeprel} in momentum space is $v {\cdot} k / v {\cdot} P \approx x$, i.e., the amplitudes $A_i$ and $B_i$ acquire an explicit $x$-dependence, which has already been pointed out in Refs. \cite{MuschThesis2,Accardi:2009au}.
For the $\Gamma$-structures at leading twist, the correlator can be written in the form
\begin{align}
\frac{1}{2 P^+}\widetilde \Phi^{[\gamma^+]}_{\text{unsubtr.}} & =
\widetilde{A}_{2B}
+ i m_N \epsilon_{ij} \vect{b}_i \vect{S}_j\, \widetilde{A}_{12B}
\displaybreak[0] \\
\frac{1}{2 P^+}\widetilde \Phi^{[\gamma^+ \gamma^5]}_{\text{unsubtr.}} & =
- \Lambda\, \widetilde{A}_{6B}
+ i \left\{ (b {\cdot} P) \Lambda - m_N (\vprp{b} {\cdot} \vprp{S}) \right\}\, \widetilde{A}_{7B}
\displaybreak[0] \\
\frac{1}{2 P^+}\widetilde \Phi^{[i \sigma^{i+} \gamma^5]}_{\text{unsubtr.}} & =
i m_N \epsilon_{ij} \vect{b}_j \, \widetilde{A}_{4B}
- \vect{S}_i\, \widetilde{A}_{9B}
- im_N \Lambda \vect{b}_i\, \widetilde{A}_{10B}
+ m_N \left\{ (b {\cdot} P) \Lambda - m_N (\vprp{b} {\cdot} \vprp{S}) \right\} \vect{b}_i\, \widetilde{A}_{11B}
\end{align}
where the indices $i,j$ correspond to transverse directions, $i,j\in \{1,2\}$
(cf.~appendix~\ref{sec-conv} for further details on notation), and where we
have introduced the following abbreviations for combinations of amplitudes:
\begin{align}
\widetilde{A}_{2B} & \ \equiv\ \widetilde{A}_2 + R({\hat \zeta}^2) \widetilde{B}_1 \nonumber \displaybreak[0] \\
\widetilde{A}_{4B} & \ \equiv\ \widetilde{A}_4 - R({\hat \zeta}^2) \widetilde{B}_3 \nonumber \displaybreak[0] \\
\widetilde{A}_{6B} & \ \equiv\ \widetilde{A}_6 + \left(1-R({\hat \zeta}^2) \right) \left\{ \widetilde{B}_{11} + R({\hat \zeta}^2) \widetilde{B}_{14} \right\} \nonumber \displaybreak[0] \\
\widetilde{A}_{7B} & \ \equiv\ \widetilde{A}_7 + R({\hat \zeta}^2) \widetilde{B}_{13} \nonumber \displaybreak[0] \\
\widetilde{A}_{9B} & \ \equiv\ \widetilde{A}_9 + R({\hat \zeta}^2) \widetilde{B}_{15} \nonumber \displaybreak[0] \\
\widetilde{A}_{10B} & \ \equiv\ \widetilde{A}_{10} - \left(1-R({\hat \zeta}^2) \right) \left\{ \widetilde{B}_{18} - R({\hat \zeta}^2) \widetilde{B}_{20} \right\} \nonumber \displaybreak[0] \\
\widetilde{A}_{11B} & \ \equiv\ \widetilde{A}_{11} - R({\hat \zeta}^2) \widetilde{B}_{17} \nonumber \displaybreak[0] \\
\widetilde{A}_{12B} & \ \equiv\ \widetilde{A}_{12} - R({\hat \zeta}^2) \widetilde{B}_8 \displaybreak[0]
\label{eq-ABamps}
\end{align}
For later convenience we also define
\begin{align}
\widetilde{A}_{9Bm} & \ \equiv\ \widetilde{A}_{9B} - \frac{1}{2} m_N^2 b^2 \widetilde{A}_{11B} \ .
\label{eq-A9m}
\end{align}
Performing the Fourier transformation and comparing with the decomposition Eqs. \eqref{eq-phigammaplus}-\eqref{eq-phisigmaplusi}, we can express the TMDs\xspace in terms of Fourier-transforms of the above amplitudes.
Using the combined amplitudes $\widetilde{A}_{iB}$,
the results are of the same form as in the straight-link case of Ref. \cite{Musch:2010ka},
\begin{align}
f_1(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 2 \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{2B} \,,\nonumber \displaybreak[0] \\
g_1(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = - 2 \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{6B} + 2 \partial_x \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{7B} \nonumber \,, \displaybreak[0] \\
g_{1T}(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 4 m_N^2 \partial_{\vprp{k}^2} \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{7B} \nonumber \,, \displaybreak[0] \\
h_1(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = - 2 \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{9Bm} \nonumber \,, \displaybreak[0] \\
h_{1L}^\perp(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 4 m_N^2 \partial_{\vprp{k}^2} \left( \ensuremath{\int_\mathcal{F}} \widetilde{A}_{10B}
+ \partial_x \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{11B} \right) \nonumber \,, \displaybreak[0] \\
h_{1T}^\perp (x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 8 m_N^{4} \left(\partial_{\vprp{k}^2}\right)^2 \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{11B}\ ,
\label{eq-tmdsfromampseven}
\end{align}
except that the abbreviation $\ensuremath{\int_\mathcal{F}}$ is now applied to the $v$-dependent amplitudes and includes the soft factor:
\begin{align}
\ensuremath{\int_\mathcal{F}} \widetilde{A}_i \equiv & \int \frac{d^2 \vprp{b}}{(2\pi)^2}\,e^{-i\vprp{b}{\cdot}\vprp{k}}\, \frac{1}{\widetilde{\mathcal{S}}(b^2;\ldots)} \int \frac{d(b {\cdot} P)}{(2\pi)}\,e^{ix(b {\cdot} P)}
\widetilde{A}_i(-\vprp{b}^2,b {\cdot} P,(b {\cdot} P) R({\hat \zeta}^2)/m_N^2,-1/(m_N{\hat \zeta})^2,\eta v {\cdot} P) \nonumber \\
= & \int_0^\infty \frac{d(-b^2)}{2(2\pi)}\ \frac{J_0(\sqrt{-b^2}\, |\vprp{k}|)}{\widetilde{\mathcal{S}}(b^2;\ldots)}\ \int \frac{d(b {\cdot} P)}{(2\pi)}\,e^{ix(b {\cdot} P)}\
\widetilde{A}_i(b^2,b {\cdot} P,(b {\cdot} P) R({\hat \zeta}^2)/m_N^2,-1/(m_N{\hat \zeta})^2,\eta v {\cdot} P) \,
\label{eq-fourint}
\end{align}
Also, there are two further TMDs\xspace that are not present in the straight-link case, the T-odd distributions
\begin{align}
f_{1T}^\perp(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 4 m_N^2 \partial_{\vprp{k}^2} \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{12B} \,,\nonumber \displaybreak[0] \\
h_1^\perp(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = - 4 m_N^2 \partial_{\vprp{k}^2} \ensuremath{\int_\mathcal{F}}\ \widetilde{A}_{4B} \, .
\label{eq-tmdsfromampsodd}
\end{align}
Again the dots ``$\ldots$'' indicate further parameters that specify the geometry of the soft factor.
The T-even distributions $f_1$, $g_1$, $h_1$, $g_{1T}$, $h_{1L}^\perp$ and $h_{1T}^\perp$ fulfill
\begin{align}
f^{\text{T-even}}(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) = f^{\text{T-even}}(x,\vprp{k}^2;{\hat \zeta},\ldots,-\eta v {\cdot} P)
\end{align}
while the T-odd distributions, i.e., at leading twist the Sivers function $f_{1T}^\perp$ and the Boer-Mulders function $h_{1}^\perp$, fulfill
\begin{align}
f^{\text{T-odd}}(x,\vprp{k}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = - f^{\text{T-odd}}(x,\vprp{k}^2;{\hat \zeta},\ldots,-\eta v {\cdot} P)
\end{align}
As a result, T-odd distributions must vanish for $\eta = 0$, which corresponds to straight gauge links.
TMDs\xspace for SIDIS and DY are obtained for $\eta v {\cdot} P \rightarrow \infty$ and $\eta v {\cdot} P \rightarrow -\infty$, respectively.
In the following, we choose $v {\cdot} P \geq 0$, such that the SIDIS and DY limits for space-like $v$ can also be written as $\eta |v| \rightarrow \infty$ and $\eta |v| \rightarrow -\infty$, respectively.
Equations \eqref{eq-tmdsfromampseven} and \eqref{eq-tmdsfromampsodd} show that certain $x$-integrated TMDs\xspace in Fourier space directly correspond to the amplitudes $\widetilde{A}_{iB}$ evaluated at $b {\cdot} P = 0$ :
\begin{align}
\tilde f_1^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 2\, \widetilde{A}_{2B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots) \,,\nonumber \displaybreak[0] \\
\tilde g_1^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = - 2\, \widetilde{A}_{6B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots) \nonumber \,, \displaybreak[0] \\
\tilde g_{1T}^{[1](1)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = -2\, \widetilde{A}_{7B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots) \nonumber \,, \displaybreak[0] \\
\tilde h_1^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = - 2\, \widetilde{A}_{9Bm}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots) \nonumber \,, \displaybreak[0] \\
\tilde h_{1L}^{\perp[1](1)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = -2\,\widetilde{A}_{10B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta
v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots) \nonumber \,, \displaybreak[0] \\
\tilde h_{1T}^{\perp[1](2)} (\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 4\,\widetilde{A}_{11B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots)\, , \nonumber \displaybreak[0] \\
\tilde f_{1T}^{\perp[1](1)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & =-2 \,\widetilde{A}_{12B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots) \,,\nonumber \displaybreak[0] \\
\tilde h_1^{\perp[1](1)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P) & = 2 \, \widetilde{A}_{4B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P)/\widetilde{\mathcal{S}}(b^2;\ldots) \, .
\label{eq-xintderfttmds}
\end{align}
The (derivatives of) Fourier-transformed TMDs\xspace $\tilde f_1^{(0)}$, $\tilde g_1^{(0)}$, $\tilde g_{1T}^{(1)}$, $\tilde h_1^{(0)}$, $\tilde h_{1L}^{\perp(1)}$, $\tilde h_{1T}^{\perp(2)}$, $\tilde f_{1T}^{\perp(1)}$ and $\tilde h_1^{\perp(1)}$
are naturally accessible from the Fourier-transformed cross section of, e.g., SIDIS \cite{Boer:2011xd}, and naturally appear in evolution equations, see, e.g. \cite{Idilbi:2004vb}.
\subsection{Generalized shifts from amplitudes}
In section \ref{sec-latquan} we have given an example that ratios of certain $\vprp{k}$-moments of TMDs\xspace have interesting physical interpretations.
These ratios, and their counterparts generalized to non-zero $\vprp{b}$, are also advantageous from a theoretical point of view:
Obviously, the soft factor $\widetilde{\mathcal{S}}$ cancels in any ratio formed from the objects in Eq. \eqref{eq-xintderfttmds}, along with any
$\Gamma$-independent multiplicative renormalization factor \cite{MuschThesis2,Hagler:2009mb,Musch:2010ka,Boer:2011xd}.
Eq. \eqref{eq-xintderfttmds} identifies $x$-integrated derivatives of Fourier-transformed TMDs\xspace with simple linear combinations of amplitudes $\widetilde{A}_i$ and $\widetilde{B}_i$ evaluated at the same values of $\vprp{b}^2$, $b \cdot P$, ${\hat \zeta}$ and $\eta v {\cdot} P$. Forming ratios of these objects thus just amounts to taking ratios of linear combinations of the fundamental correlators $\tilde \Phi^{[\Gamma]}_\text{unsubtr.}$ evaluated at the same point, i.e., with the same values for $b$, $P$ and $\eta v$. For a discussion of the renormalization properties of ratios of the objects in \eqref{eq-xintderfttmds} it is thus sufficient to
understand the renormalization properties of (ratios formed from) the
correlators $\widetilde \Phi^{[\Gamma]}_\text{unsubtr.} = \frac{1}{2} \, \bra{P,S}\ \bar{q}(0)\, \Gamma\ \mathcal{U}\ q(b)\ \ket{P,S} $.
Analytical studies of the operator $\bar{q}(0)\, \Gamma\, \mathcal{U} q(b)$ in the continuum \cite{Dotsenko:1979wb,Craigie:1980qs,Arefeva:1980zd,Aoyama:1981ev,Stefanis:1983ke,Dorn:1986dt}
suggest that for $b^2 \neq 0$ the renormalization factors are multiplicative and $\Gamma$-independent. The basic reason is that the quark field operators are at different locations and undergo wave function renormalization separately. We will assume here that our lattice representation of $\bar{q}(0)\, \Gamma\, \mathcal{U}\, q(b)$ is renormalized multiplicatively independent of $\Gamma$ as long as we keep $\vprp{b}^2$ larger than a few lattice spacings. A more detailed discussion and numerical studies of the renormalization properties of this operator can be found in Ref. \cite{Musch:2010ka}. It remains an interesting task for the future to perform a more thorough treatment of non-local operators on the lattice. Under the assumption of multiplicative renormalization, generalized shifts such as $\langle \vect{k}_y \rangle_{TU}(\vprp{b}^2;{\hat \zeta},\eta v {\cdot} P) \equiv m_N \tilde f_{1T}^{\perp[1](1)} / \tilde f_1^{[1](0)}$ can only depend on $\vprp{b}^2$, ${\hat \zeta}$ and on the staple extent $\eta v {\cdot} P$. All other renormalization and soft factor related dependences cancel out in the ratio.
In this work, we will present numerical results for the following generalized shifts:
\begin{align}
\langle \vect{k}_y \rangle_{TU}(\vprp{b}^2;{\hat \zeta},\eta v {\cdot} P)
&\ \equiv\ m_N \frac{\tilde f_{1T}^{\perp[1](1)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)}{\tilde f_1^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)}
= - m_N \frac{\widetilde{A}_{12B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) }{ \widetilde{A}_{2B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) } \nonumber \\
&\ \xrightarrow{\vprp{b}^2 = 0}
\left. \frac{ \int dx \int d^2 \vprp{k}\, \vect{k}_y \ \Phi^{[\gamma^+]}(x,\vprp{k},P,S;\ldots) }
{ \int dx \int d^2 \vprp{k} \phantom{\vect{k}_y}\ \Phi^{[\gamma^+]}(x,\vprp{k},P,S;\ldots) } \right|_{\displaystyle \vprp{S}=(1,0) }
\label{eq-genSivShift} \displaybreak[0] \\
\langle \vect{k}_y \rangle_{UT}(\vprp{b}^2;{\hat \zeta},\eta v {\cdot} P)
&\ \equiv\ m_N \frac{\tilde h_{1}^{\perp[1](1)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)}{\tilde f_1^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)}
= m_N \frac{\widetilde{A}_{4B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) }{ \widetilde{A}_{2B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) } \nonumber \\
&\ \xrightarrow{\vprp{b}^2 = 0}
\left. \frac{ \sum_{\Lambda=\pm 1} \int dx \int d^2 \vprp{k}\, \vect{k}_y \ \Phi^{[\gamma^+ + s^j i \sigma^{j+}\gamma^5]}(x,\vprp{k},P,S;\ldots) }
{ \sum_{\Lambda=\pm 1} \int dx \int d^2 \vprp{k} \phantom{\vect{k}_y}\ \Phi^{[\gamma^+ + s^j i \sigma^{j+}\gamma^5]}(x,\vprp{k},P,S;\ldots) } \right|_{\displaystyle \vprp{s}=(1,0) }
\label{eq-genBMShift} \displaybreak[0] \\
\langle \vect{k}_x \rangle_{TL}(\vprp{b}^2;{\hat \zeta},\eta v {\cdot} P)
&\ \equiv\ m_N \frac{\tilde g_{1T}^{[1](1)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)}{\tilde f_1^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)} =
- m_N \frac{\widetilde{A}_{7B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) }{ \widetilde{A}_{2B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) } \nonumber \\
&\ \xrightarrow{\vprp{b}^2 = 0}
\left. \frac{ \int dx \int d^2 \vprp{k}\, \vect{k}_x \ \Phi^{[\gamma^+ + \lambda \gamma^{+}\gamma^5]}(x,\vprp{k},P,S;\ldots) }
{ \int dx \int d^2 \vprp{k} \phantom{\vect{k}_x}\ \Phi^{[\gamma^+ + \lambda \gamma^{+}\gamma^5]}(x,\vprp{k},P,S;\ldots) } \right|_{\displaystyle \vprp{S}=(1,0),\ \lambda=1 }
\label{eq-geng1TShift} \, ,\\
\frac{\tilde h_{1}^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)}{\tilde f_1^{[1](0)}(\vprp{b}^2;{\hat \zeta},\ldots,\eta v {\cdot} P)}
&= - \frac{\widetilde{A}_{9Bm}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) }{ \widetilde{A}_{2B}(-\vprp{b}^2,0,0,-1/(m_N{\hat \zeta})^2,\eta v{\cdot} P) }
\nonumber \\
&\ \xrightarrow{\vprp{b}^2 = 0}
\left.
\frac{ \int dx \int d^2 \vprp{k}\, \ \Phi^{[s^j i \sigma^{j+}\gamma^5]}(x,\vprp{k},P,S;\ldots) }
{ \int dx \int d^2 \vprp{k} \ \Phi^{[\gamma^+]}(x,\vprp{k},P,S;\ldots) } \right|_{\displaystyle \vprp{S}=(1,0),\ \vprp{s}=(1,0) }
\label{eq-h1overf1} \, .
\end{align}
\begin{itemize}
\item
The ``generalized Sivers shift'' $\langle \vect{k}_y \rangle^{\text{Sivers}}=\langle \vect{k}_y \rangle_{TU}$ has already been discussed in section \ref{sec-latquan}.
It is T-odd, i.e., we expect to obtain results of opposite sign in the SIDIS and DY limits $\eta v {\cdot} P \rightarrow \infty$ and $\eta v {\cdot} P \rightarrow -\infty$, respectively.
The generalized Sivers shift describes a feature of the transverse momentum distribution of (unpolarized) quarks in a transversely polarized proton.
In the formal limit $\vprp{b}^2 = 0$ it measures the dipole moment of that distribution orthogonal to the polarization of the proton.
\item
The ``generalized Boer-Mulders shift'' $\langle \vect{k}_y \rangle^{\text{BM}}=\langle \vect{k}_y \rangle_{UT}$ is also T-odd and addresses the distribution of transversely polarized quarks in an unpolarized proton.
In the limit $\vprp{b}^2 = 0$, the Boer-Mulders shift describes the dipole moment of that distribution orthogonal to the polarization of the quarks.
Note that we use a sum over proton helicities $\sum_{\Lambda=\pm 1}$ in Eq. \eqref{eq-genBMShift} to represent the unpolarized target nucleon.
\item
The generalized shift $\langle \vect{k}_x \rangle^{{g_{1T}}}=\langle \vect{k}_x \rangle_{TL}$
attributed to the ``worm gear'' function $g_{1T}$
quantifies a dipole deformation of the transverse momentum distribution induced by the correlation of the quark helicity and the transverse proton spin.
Unlike the Sivers and the Boer-Mulders shifts, it is a T-even quantity, i.e., the SIDIS and DY limits $\eta v {\cdot} P \rightarrow \pm \infty$ are expected to be the same.
This shift has already been studied in lattice QCD using straight gauge links \cite{MuschThesis2,Hagler:2009mb,Musch:2010ka}.
We are interested to see by how much this ``process independent'' result obtained at $\eta=0$
differs from the results calculated with SIDIS- and DY-type gauge links in the limit $\eta v{\cdot} P \rightarrow \pm \infty$.
\item
The ratio $\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)}$ can be identified with a ``generalized tensor charge".
Clearly, it is also a T-even quantity, i.e., no differences are expected
between the SIDIS and DY limits $\eta v {\cdot} P \rightarrow \pm \infty$.
We have studied $\tilde h_{1}^{[1](0)}$ already in \cite{MuschThesis2,Hagler:2009mb,Musch:2010ka} on the lattice using straight gauge links.
As $\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)}$ doesn't involve any $\vect{k}$-weighting and is directly related to the well-known
transversity and unpolarized distribution functions, we expect it to be a particularly clean observable.
It therefore qualifies as a very good candidate for our study of the $\eta|v|$-dependence of T-even observables, in particular
the transition from straight to staple-shaped gauge links.
\end{itemize}
The framework laid out above provides the basis for our numerical lattice
calculations described in the next section. Before proceeding, it is
worth reiterating the logic underlying our approach. Recognizing that
the generic kinematics for which TMDs\xspace are defined are space-like, with
light-like separations representing a special limiting case, we proceed
by considering kinematics off the light cone from the start. We again
emphasize that, whereas we thoroughly examine the behavior of our data
as the kinematics are pushed in the direction of the light cone, statements
about formal properties of the light-cone limit lie beyond the
purview of this investigation. Having parametrized the relevant nonlocal
matrix element in terms of Lorentz-invariant amplitudes,
cf.~section~\ref{sec-parametr}, we choose to perform its evaluation
in a Lorentz frame in which the operator under consideration is defined
at one fixed time. There is no obstacle to this choice in view of the
space-like separations entering the original definition of the matrix
element. In this frame, we cast the computation of the matrix element in
terms of a Euclidean path integral, which we evaluate employing lattice
QCD, as detailed in the next section.
\section{Lattice Calculations}
\label{sec-latticecalc}
\subsection{Simulation setup and parameters}
The methodology we use to calculate the non-local correlators on the lattice has been described in detail in Ref. \cite{Musch:2010ka}, except that we now extend this method to staple-shaped links.
Again, we employ MILC lattices \cite{Ber01,Aubin:2004wf} that have been previously used by the LHP collaboration for GPD calculations \cite{Hagler:2007xi}; however, compared to our previous work with straight gauge links, we now go to lighter pion masses and make use of the coherent proton and anti-proton sequential propagators of Ref. \cite{Bratt:2010jn} to increase our statistics.
The new LHPC data set offers forward propagators at four different source locations on each gauge configuration. Moreover, coherent proton and antiproton sequential propagators have been calculated, each one implementing simultaneously four nucleon sink locations per gauge configuration. This way it is possible to conduct eight measurements of a three-point function on each gauge configuration in well separated areas of the lattice, boosting statistics significantly. The source-sink separation has been chosen to be nine lattice units. The simulation parameters are summarized in Table \ref{tab-gaugeconfs}.
\begin{table*}[tbp]
\centering
\renewcommand{\arraystretch}{1.1}
\begin{tabular}{|ll|l|c||c|c|c||c|c|}
\hline
$\hat m_{u,d}$ & $\hat m_{s}$ & $\hat L^3 \times \hat T$\rule{0ex}{1.2em} & $10/g^2$ & $a\units{(fm)}$ & $m_\pi^\text{DWF}\units{(MeV)}$ & $m_N^\text{DWF}\units{(GeV)}$ & $\#$conf. & $\#$meas. \\
\hline
$0.01$ & $0.05$ & $28^3 \times 64$ & $6.76$ & $0.11967(14)(99)$ & $369.0(09)(35)$ & 1.197(09)(12) & 273 & 2184 \\
$0.01$ & $0.05$ & $20^3 \times 64$ & $6.76$ & $0.11967(14)(99)$ & $369.0(09)(35)$ & 1.197(09)(12) & 658 & 5264 \\
$0.02$ & $0.05$ & $20^3 \times 64$ & $6.79$ & $0.11849(14)(99)$ & $518.4(07)(49)$ & 1.348(09)(13) & 486 & 3888 \\
\hline
\end{tabular}\par\vspace{1ex}
\renewcommand{\arraystretch}{1.0}
\caption{Lattice parameters of the $n_f = 2{+}1$ MILC gauge configurations \cite{Ber01,Aubin:2004wf} used in this work. The lattice spacing $a$ has been obtained from the ``smoothed'' values for $r_1/a$ given in Ref. \cite{Bazavov:2009bb} and the value $r_1= 0.3133(26)\units{fm}$ from the analysis of Ref.~\cite{Davies:2009tsa}. The first error estimates statistical errors in $r_1/a$, the second error originates from the uncertainty about $r_1$ in physical units. We also list the pion and the nucleon masses determined in Ref. \cite{Bratt:2010jn} with the LHPC propagators using domain wall valence fermions. The first error is statistical, the second error comes from the conversion to physical units using $a$ as quoted in the table. Note that the masses quoted here in physical units differ slightly from those listed in Refs. \cite{Hagler:2007xi,Bratt:2010jn}, because these references use a different scheme to fix the lattice spacing. The second to last column lists the number of gauge configurations and the last column shows the resulting number of measurements for the calculation of three-point functions achieved by means of multiple locations for source and sink.
\label{tab-gaugeconfs}%
}
\end{table*}
\subsection{Nucleon momenta, choice of link directions, and extraction of
amplitudes}
For all of the ensembles listed in Table~\ref{tab-gaugeconfs}, nucleon
momenta $\vect{P} =0$ and $\vect{P} =2\pi/(a\hat{L} ) \cdot (-1,0,0)$,
implemented via corresponding momentum projections in the sequential
propagators, were available. In addition, sequential propagators were
produced corresponding to the nucleon momenta
$\vect{P} =2\pi/(a\hat{L} ) \cdot (-2,0,0)$ and
$\vect{P} =2\pi/(a\hat{L} ) \cdot (1,-1,0)$ for the $\hat{m}_{u,d} =0.02$
ensemble only. We extracted the matrix element
$\widetilde \Phi_\text{unsubtr.}^{[\Gamma]}(b,P,S;\mathcal{C}_b) \equiv
\frac{1}{2}\, \bra{P,S}\ \bar{q}(0)\,
\Gamma\ \WlineC{\mathcal{C}_b}\ q(b)\ \ket{P,S} $ from plateaux
in standard three-point function to two-point function ratios, for a
complete basis of $\Gamma $ structures and nucleon states polarized in the
3-direction. The nucleon momenta $\vect{P} $, quark separations $\vect{b} $
and corresponding staple-shaped gauge link paths $\mathcal{C}_b $ used
on the lattice in the present investigation are listed in
Table~\ref{stapletable}. The link path $\mathcal{C}_b $ is characterized
by the quark separation vector $\vect{b} $ and the staple vector
$\eta \vect{v} $, cf.~Fig.~\ref{fig-links}. The range of $\eta $ studied
was always chosen to extend from zero to well beyond the point where a
numerical signal ceases to be discernible. Furthermore,
it should be noted that in the case of either $\vect{b} $ or $\vect{v} $
extending into a direction in a lattice plane which forms an angle of
$\pi/4$ with the lattice axes spanning the plane, there are two optimal
approximations of the corresponding continuum path by a lattice link path;
e.g., if one denotes the lattice link vector in $i$-direction as
$\vect{e}_{i} $, then $\vect{b} = 2 (\vect{e}_{1} +\vect{e}_{2} ) $ is
equally well approximated by the sequence of links
$(\vect{e}_{1},\vect{e}_{2},\vect{e}_{1},\vect{e}_{2})$ as by the sequence
$(\vect{e}_{2},\vect{e}_{1},\vect{e}_{2},\vect{e}_{1})$. As far as
$\vect{b} $ is concerned, in such a situation, our calculations
always included both optimal link paths. However, in the case of
$\vect{v} $, in these situations, only one of the two link paths was
included. To be specific, in the instances of
$\eta \vect{v} =\pm n^{\prime } (\vect{e}_{1} \pm \vect{e}_{i} )$ quoted
in Table~\ref{stapletable}, the link path always departs from the quark
locations in $i$-direction, not $1$-direction. This is a shortcoming
of the discretization which breaks the manifest T-transformation
properties present for the continuum staple; presumably it is
responsible for the problematic mixing of T-even and T-odd amplitudes
which we observe in our analysis in the case of staple directions
off the lattice axes. While we expect a symmetry-improved calculation
including both optimal link paths to avoid this issue, with the presently
available data, we find that we need to impose explicitly T-odd/T-even symmetry
in the system of equations from which we extract the amplitudes whenever
$\vect{v} $ does not coincide with a lattice axis.
\begin{table}
\begin{tabular}{|c|c|r|c|}
\hline
$\vect{b} /a $ & $\eta \vect{v} /a $ & $\vect{P} \cdot a\hat{L} /(2\pi )$ &
Notes \\
\hline\hline
$n\cdot (0,0,1), n=-7,\ldots,7$ & $\pm n^{\prime } \cdot (1,0,0)$ &
$(0,0,0)$ & \\
\cline{3-4}
& & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{2-4}
& $\pm n^{\prime } \cdot (1,1,0)$ & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{2-4}
& $\pm n^{\prime } \cdot (1,0,0)$ & $(1,-1,0)$ & $\hat{m}_{u,d} =0.02$
ensemble only \\
\cline{2-4}
& $\pm n^{\prime } \cdot (1,-1,0)$ &
$(1,-1,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{1-4}
$n\cdot (0,1,0), n=-7,\ldots,7$ & $\pm n^{\prime } \cdot (1,0,0)$ &
$(0,0,0)$ & \\
\cline{3-4}
& & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{2-4}
& $\pm n^{\prime } \cdot (0,0,1)$ & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{2-4}
& $\pm n^{\prime } \cdot (1,0,1)$ & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{1-4}
$n\cdot (0,1,1), n=-2,\ldots,2$ & $\pm n^{\prime } \cdot (1,0,0)$ &
$(0,0,0)$ & \\
\cline{3-4}
& & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{1-4}
$n\cdot (0,-1,1), n=-2,\ldots,2$ & $\pm n^{\prime } \cdot (1,0,0)$ &
$(0,0,0)$ & \\
\cline{3-4}
& & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{1-4}
$\pm (0,3,\pm 2)$ & $\pm n^{\prime } \cdot (1,0,0)$ & $(0,0,0)$ & \\
\cline{3-4}
& & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{1-4}
$\pm (0,4,\pm 2)$ & $\pm n^{\prime } \cdot (1,0,0)$ & $(0,0,0)$ & \\
\cline{3-4}
& & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{1-4}
$\pm (0,4,\pm 3)$ & $\pm n^{\prime } \cdot (1,0,0)$ & $(0,0,0)$ & \\
\cline{3-4}
& & $(-1,0,0)$ & \\
\cline{3-4}
& & $(-2,0,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\cline{1-4}
$n\cdot (1,1,0), n=-4,\ldots,4$ &
$\pm n^{\prime } \cdot (1,-1,0)$ &
$(1,-1,0)$ & $\hat{m}_{u,d} =0.02$ ensemble only \\
\hline
\end{tabular}
\caption{Sets of staple-shaped gauge link paths and nucleon momenta
$\vect{P} $ used on the lattice. Gauge link paths are characterized by the
quark separation vector $\vect{b} $ and the staple vector $\eta \vect{v}$,
cf.~Fig.~\ref{fig-links}. The surveyed range of $\eta $, parameterized in
the table by the integer $n^{\prime } $, was always chosen to extend from
zero to well beyond the point where a numerical signal ceases to be
discernible. The maximal magnitude of the Collins-Soper parameter
$\hat{\zeta} $ attained in these sets is $|\hat{\zeta} |=0.78$, for
$\vect{P}\cdot a\hat{L}/(2\pi)=(-2,0,0)$ paired with
$\eta \vect{v}/a =\pm n^{\prime } \cdot (1,0,0)$.}
\label{stapletable}
\end{table}
In practice, the overdetermined system of equations which we solve in order
to relate the matrix elements $\widetilde \Phi_\text{unsubtr.}^{[\Gamma]} $ to
the corresponding amplitudes is set up in terms of the quantities
$\tilde{a}_{i} , \tilde{b}_{i} $, cf.~Eq.~(\ref{eq-ab}) in conjunction
with Eqs.~(\ref{eq-phitildedecomp-scal})-(\ref{eq-phitildedecomp-tens}).
This form is suited to include the case $\hat \zeta=0$, where the sign of the
prefactor in front of $\tilde{b}_{i} $ depends on whether the limit
$\eta v {\cdot} P = 0$ is approached from the SIDIS or the DY side.
\subsection{Numerical Results}
\subsubsection{The generalized Sivers shift}
In the following, we concentrate on results for the isovector, $u-d$ quark
combination, because in this case contributions from disconnected diagrams and possible vacuum expectation values cancel out.
The errors shown are statistical only. At the present level of accuracy in this exploratory study, we set aside a quantitative analysis of systematic errors. We use the central values for the lattice spacing $a$ as given in Table \ref{tab-gaugeconfs} to convert to physical units. For $m_N$, we consistently substitute the value of the nucleon mass as determined on the lattice, rather than the physical nucleon mass.
Figures \ref{fig-Sivers_etadepend} to \ref{fig-Sivers_evolution_combined} show our results for the generalized Sivers shift, $\langle \vect{k}_y \rangle_{u-d}^{\text{Sivers}}$.
We begin with a discussion of its dependence on the staple orientation, i.e., SIDIS- or DY-like, and the staple extent, $\eta |v|$,
as displayed in Fig.~\ref{fig-Sivers_etadepend} for a Collins-Soper evolution parameter of
$\hat \zeta = 0.39$ and a pion mass of $m_\pi = 518 \units{MeV}$.
As mentioned before, the T-odd Sivers function must vanish for $\eta |v|=0$, i.e., a straight Wilson line between the quark fields,
but non-vanishing results are allowed (and generally expected) for non-zero staple extents.
Furthermore, the T-odd observables are anti-symmetric in $\eta |v|$, so we expect the Sivers shift to be of the same size but
opposite in sign for the SIDIS and the DY cases.
This is exactly what we find in, e.g., Fig.~\ref{fig-Sivers_lsqr-1_zetasqrlat4}, showing the shift for a quark-antiquark distance of
a single lattice spacing, $|\vprp{b}|=1a$.
The aforementioned features are realized in form of a curve that is reminiscent of a hyperbolic tangent.
We stress that the observed zero crossing with a change in sign is directly caused by the underlying gauge-invariant operators and their
symmetry properties, and hence represents a consistency check of our calculation rather than any sort of a prediction.
Remarkably, already as $|\eta| |v|$ approaches values of $\sim 6a$, we find that the Sivers shift stabilizes and reaches specific plateau values.
Apart from finite volume effects, in particular wrap-around effects due to the periodic boundary conditions on the lattice,
we see no reason to expect that once a plateau has been reached,
the value of the shift would significantly change as $|\eta| |v|\rightarrow\infty$.
To obtain first estimates for staple-shaped Wilson lines that have an infinite extent in $v$-direction,
we therefore choose to average the shifts in the plateau regions $|\eta| |v|=7a \ldots 12a$, as illustrated by the straight lines.
Clearly, as $|\vprp{b}|$ increases from $0.12\units{fm}$ in Fig.~\ref{fig-Sivers_lsqr-1_zetasqrlat4} to
$0.47\units{fm}$ in Fig.~\ref{fig-Sivers_lsqr-16_zetasqrlat4}, the signal-to-noise ratio decreases as we approach larger values of $|\eta| |v|$.
For smaller $|\eta| |v|$
the statistical uncertainties are much smaller, and the corresponding values tend to dominate the averages when
the errors are taken into account as weights.
At the same time, however, these statistically dominating data points are more likely to introduce systematic uncertainties related to the (unknown) onset of the ``true" plateau region and the corresponding starting value for the averaging procedure.
Therefore, in order to avoid a too strong bias from the data at smaller $|\eta| |v|$,
we do not use the respective statistical errors as weights in the averaging.
Our final estimates for the Sivers shift are obtained from the mean value of the SIDIS and DY averages and by imposing antisymmetry
in $\eta |v|$. The results are displayed as open diamonds at $\eta |v|=\pm\infty$ in Fig.~\ref{fig-Sivers_etadepend}.
The dependence of these results on $|\vprp{b}|$ is shown in Fig.~\ref{fig-Sivers_bdepend}.
In summary, for $\hat \zeta = 0.39$ and $|\vprp{b}|=0.12\ldots0.47\units{fm}$, we find a sizeable negative Sivers shift for $u-d$ quarks
in the range of $\langle \vect{k}_y \rangle_{u-d}^{\text{Sivers,SIDIS}}=-0.3\ldots-0.15 \units{GeV}$.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-Sivers_lsqr-1_zetasqrlat4}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_Sivers_lsqr-1_zetasqrlat4}
}\hfill%
\subfloat[][]{%
\label{fig-Sivers_lsqr-4_zetasqrlat4
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_Sivers_lsqr-4_zetasqrlat4}
}\\%
\subfloat[][]{%
\label{fig-Sivers_lsqr-9_zetasqrlat4}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_Sivers_lsqr-9_zetasqrlat4}
}\hfill%
\subfloat[][]{%
\label{fig-Sivers_lsqr-16_zetasqrlat4
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_Sivers_lsqr-16_zetasqrlat4}
\caption[SIDIS diagram]{%
Extraction of the generalized Sivers shift on the lattice with $m_\pi = 518 \units{MeV}$
using a lattice nucleon momentum $|\vect{P}^{\text{lat}}| = 2\pi /(a\hat{L}) \approx 500 \units{MeV}$
at the corresponding maximal Collins-Soper evolution parameter $\hat \zeta = 0.39$.
The continuous horizontal lines are obtained from two independent averages of the data points
with staple extents in the ranges $\eta |v| = 7a .. 12a$ and $\eta |v| = -12a .. -7a$, respectively.
The outer data points shown with empty symbols have been obtained from an anti-symmetrized mean value of these averages,
i.e., the expected T-odd behavior of the Sivers shift has been put in explicitly. These outer
data points are our estimates for the asymptotic values at $\eta |v| \rightarrow \pm \infty$ and thus
represent the generalized Sivers shifts for SIDIS and DY.
Error bars show statistical uncertainties only.
Figures \subref{fig-Sivers_lsqr-1_zetasqrlat4} and \subref{fig-Sivers_lsqr-4_zetasqrlat4}
have been obtained with rather small quark field separations $|\vprp{b}|=1a$ and $2a$.
Therefore, they might be affected by significant lattice cutoff effects.
\label{fig-Sivers_etadepend}%
}
\end{figure}
\begin{figure}[btp]
\centering%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_SiversRat_zetasqrlat4_bdepend}
\caption{%
Generalized Sivers shift as a function of the quark separation $|\vprp{b}|$
for the SIDIS case ($\eta |v| = \infty$), extracted
on the lattice with $m_\pi = 518 \units{MeV}$ for $\hat \zeta = 0.39$.
The data points lying in the shaded area below $|\vprp{b}| \approx 0.25 \units{fm}$ might be affected by significant lattice cutoff effects.
Error bars show statistical uncertainties only.
\label{fig-Sivers_bdepend}%
}
\end{figure}
Next, we turn to the dependence of our results on the Collins-Soper evolution parameter $\hat\zeta$.
In Fig.~\ref{fig-Sivers_zetasqrlat_0_16}, we consider two ``extreme" cases, namely, a vanishing $\hat\zeta$ as well as the largest
$\hat\zeta=0.78$ that we could access in this study.
While we find rather precise values for the Sivers shift for $\hat\zeta=0$ with a well-defined plateau\footnote{Note that, in the case at hand, $\hat\zeta=0$
corresponds to $\vect{P}=0$, so that one cannot identify a ``forward'' or ``backward'' direction.
Hence, there is only a single branch in $\eta |v|$, the sign of which is a matter of definition, see also the discussion further below in the text.}
for $|\eta| |v|\ge 6a$ in Fig.~\ref{fig-Sivers_lsqr-9_zetasqrlat0}, fluctuations and uncertainties
quickly increase with $|\eta||v|$ for $\hat\zeta=0.78$ in Fig.~\ref{fig-Sivers_lsqr-9_zetasqrlat16}.
In particular, it is difficult to identify the onset of a plateau on the right hand (SIDIS) side of Fig.~\ref{fig-Sivers_lsqr-9_zetasqrlat16}.
Following the averaging procedure described above, we however find that the estimated values at $|\eta| |v|=\infty$ for the two extreme cases of $\hat\zeta$
agree within uncertainties.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-Sivers_lsqr-9_zetasqrlat0}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_Sivers_lsqr-9_zetasqrlat0}
}\hfill%
\subfloat[][]{%
\label{fig-Sivers_lsqr-9_zetasqrlat16
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_Sivers_lsqr-9_zetasqrlat16}
}\\%
\caption[SIDIS diagram]{%
Generalized Sivers shift on the lattice with $m_\pi = 518 \units{MeV}$
for a quark separation of three lattice spacings, $|\vprp{b}| = 3a = 0.36 \units{fm}$, extracted at $\hat \zeta = 0$
and at our highest value of the Collins-Soper evolution parameter, $\hat \zeta = 0.78$.
Figure \subref{fig-Sivers_lsqr-9_zetasqrlat16} has been obtained from nucleons with momentum $|\vect{P}^{\text{lat}}| = 2 \times 2\pi /(a\hat{L} ) \approx 1 \units{GeV}$ on the lattice. Error bars show statistical uncertainties only.
\label{fig-Sivers_zetasqrlat_0_16}%
}
\end{figure}
Figure \ref{fig-Sivers_evolution} shows the Sivers shift as a function of $\hat\zeta$, for $|\vprp{b}| =0.36 \units{fm}$
and a pion mass of $m_\pi = 518 \units{MeV}$.
Within the present uncertainties, we observe a statistically significant negative shift; however, it is not possible to
identify a clear trend of the data points as $\hat\zeta$ increases.
With respect to data points obtained for staple link directions $\vect{v}$ off the lattice axes,
i.e., $\hat \zeta\approx 0.55$ in Fig.~\ref{fig-Sivers_evolution},
we note again that we need to impose the T-odd/T-even (anti-)symmetry already when we solve our system of equations,
in order to avoid problematic mixings of T-even and T-odd amplitudes.
As already mentioned further above, we
expect this to become unnecessary in the case that lattice symmetry improved operators (see Appendix D of
Ref.~\cite{Musch:2010ka}) are used.
We find it very interesting to note that the contribution from $\tilde A_{12}$ alone in the numerator of Eq. \eqref{eq-genSivShift} (rather than $\tilde A_{12B}$), illustrated by the open squares,
is essentially compatible with zero within errors for all accessible values of $\hat \zeta$.
The main contribution to the transverse shift therefore comes from $-R(\hat \zeta^2)\tilde B_{8}=-\eta\,(v\cdot P)\,R(\hat \zeta^2)\tilde b_8$
(see Eqs.~(\ref{eq-ab})), i.e., the amplitude $\tilde b_8$.
Note again that, on the lattice, we employ expressions in terms of the lower-case $\tilde a_i$ and $\tilde b_i$ amplitudes, e.g. $\tilde b_8$, as they are well defined even when $\hat \zeta\rightarrow0$.
In this limit, $v\cdot P\rightarrow0$, and hence the prefactor behaves as $-(v\cdot P)\,R(\hat \zeta^2)\rightarrow a m_N$.
The sign of the prefactor of $\tilde b_i$ depends on whether one approaches the limit $v\cdot P\rightarrow0$ from the SIDIS or the DY side.
An example that explicitly shows the relative smallness of $\tilde A_{12}$ is given in Fig.~\ref{fig-Sivers_acontrib},
for $\hat \zeta = 0.39$ and $|\vprp{b}| =0.36 \units{fm}$.
While $\tilde A_{12}$ as a function of $\eta |v|$ shows the typical behavior expected for a T-odd amplitude,
it represents only about $10\%$ of the total contribution for, e.g., $|\eta||v|=6a$.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-Sivers_evolution
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_SiversRat_lsqr-9_evolution}
}\hfill%
\subfloat[][]{%
\label{fig-Sivers_acontrib}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_Sivers_lsqr-9_zetasqrlat4_acontrib}
}\\%
\caption[SIDIS diagram]{%
Generalized Sivers shift on the lattice with $m_\pi = 518 \units{MeV}$ for a quark separation of three lattice spacings, $|\vprp{b}| = 3a =0.36 \units{fm}$.
In Figure \subref{fig-Sivers_evolution} we show the $\hat \zeta$-dependence of the generalized Sivers shift, depicting both the full result and the result obtained
with just $\widetilde{A}_{12}$ in the numerator. The data points correspond to those displayed in the SIDIS limit $\eta |v| \rightarrow \infty$ in plots such as Fig. \subref{fig-Sivers_acontrib}.
Figure \subref{fig-Sivers_acontrib} shows the $\eta$-dependence at $\hat \zeta = 0.39$ for both the full result (diamonds)
and the contribution from amplitude $\widetilde{A}_{12}$ in the numerator (squares).
Asymptotic results corresponding to SIDIS and DY have been extracted as in Fig. \ref{fig-Sivers_etadepend}.
Error bars show statistical uncertainties only.
\label{fig-Sivers-aevolution}%
}
\end{figure}
As one of our central results, we show in Fig.~\ref{fig-Sivers_evolution_combined}
the Sivers shift as a function of $\hat \zeta$ for all considered ensembles, as before for a fixed $|\vprp{b}| =0.36 \units{fm}$.
Within statistical uncertainties, the data points for the two different pion masses $m_\pi = 369 \units{MeV}$ and $m_\pi = 518 \units{MeV}$,
as well as the spatial lattice volumes $V\approx(2.4 \units{fm})^3$ and $V\approx(3.4 \units{fm})^3$, are overall well compatible.
Apart from the less well determined data point at $\hat\zeta\approx 0.55$, we find
a clearly non-zero negative Sivers shift in the range
$\langle \vect{k}_y \rangle_{u-d}^{\text{Sivers,SIDIS}}=-0.48\ldots-0.2 \units{GeV}$.
Together with the relatively mild $\vprp{b}$-dependence at smaller $\vprp{b}$, cf.~Fig.~\ref{fig-Sivers_bdepend},
this provides strong evidence that the ($x$- and $\vprp{k}$-moment of the)
Sivers function $f_{1T}^{\perp}$ considered here is sizeable and negative
for $u-d$ quarks.
Our preliminary separate data for $u$- and for $d$-quarks (not shown in this work) furthermore indicate that
$f_{1T}^{\perp,u}<0$ and $f_{1T}^{\perp,d}>0$.
Although our results for the T-odd Sivers effect are still subject to many systematic effects and uncertainties, it is
interesting to note that they are overall well compatible with results from a phenomenological analysis of SIDIS data
\cite{Anselmino:2005ea,Anselmino:2011gs}, as well as arguments based on the chromodynamic lensing mechanism by Burkardt \cite{Burkardt:2002ks,Burkardt:2003je,Burkardt:2003uw}.
It should also be noted that, in a recent twist-3 analysis of single spin asymmetries from RHIC experiments,
a possible discrepancy has been found with respect to the signs \cite{Kang:2011hk}.
We stress again that fully quantitative predictions for, or a comparison with,
phenomenological and experimental TMD studies employing QCD factorization
would require lattice data for much larger Collins-Soper parameters,
$\hat\zeta \gg 1$.
With $\hat \zeta^2=- (v {\cdot} P)^2/(v^2m_N^2)$, the limit $\hat\zeta\rightarrow\infty$
corresponds to the limit of a light-like staple direction $v$,
or an infinite rapidity $y_v\rightarrow -\infty$ in Eq.~(\ref{eq-zeta}).
For the shifts and ratios defined in Eqs.~(\ref{eq-genSivShift})-(\ref{eq-h1overf1}),
where the soft factors in the TMD definitions \cite{Ji:2004wu,Aybat:2011zv,CollinsBook2011}
cancel out, large values of $\hat\zeta$ can also be accessed through large nucleon momenta.
Clearly, the limit of an infinite Collins-Soper parameter is in practice not accessible on the lattice, so that
we have to rely on results for a limited range of $\hat \zeta$, as for example in Fig.~\ref{fig-Sivers_evolution_combined}.
From the perturbative prediction for the $\zeta$ dependence, cf., e.g., Ref.~\cite{Aybat:2011zv}, we would expect that
the ratios of TMDs should become independent of $\hat\zeta$ as $\hat\zeta\rightarrow\infty$.
It would be very interesting to investigate this on the basis of future lattice results for
larger hadron momenta and with substantially improved statistics.
\begin{figure}[btp]
\centering%
\includegraphics[scale=0.9]{plots/UminusD_SiversRat_lsqr-9_evolution_combined_v2}
\caption{%
Comparison of the $\hat \zeta$-evolution of the generalized Sivers shift at $|\vprp{b}|=3a=0.36\units{fm}$
for the three different lattices listed in Table \ref{tab-gaugeconfs}.
Filled symbols correspond to the full SIDIS result. The data points with open symbols have been
obtained with only $\widetilde{A}_{12}$ in the numerator. Error bars show statistical uncertainties only.
\label{fig-Sivers_evolution_combined}%
}
\end{figure}
\subsubsection{The generalized Boer-Mulders shift}
We now turn to the second prominent T-odd TMD, the Boer-Mulders function.
Our results for the generalized Boer-Mulders shift $\langle \vect{k}_y \rangle_{u-d}^{\text{BM}}$ (Eq.~(\ref{eq-genBMShift}))
are summarized in Figs.~\ref{fig-BoerMuld_lsqr-9_zetasqrlat4} to \ref{fig-BoerMuld_evolution_combined}.
A typical example for the $\eta |v|$-dependence is shown in Fig.~\ref{fig-BoerMuld_lsqr-9_zetasqrlat4}
for a pion mass of $m_\pi = 518 \units{MeV}$, $\hat \zeta = 0.39$, and $|\vprp{b}|=0.36\units{fm}$.
Apart from the magnitude of the shift, the results are very similar to what we have found for the Sivers shift
in Fig.~\ref{fig-Sivers_lsqr-9_zetasqrlat4} above, with indications for plateaus for $|\eta| |v|\ge 6a$.
Figure \ref{fig-BoerMuld_bdepend} illustrates the dependence on $|\vprp{b}|$ for the SIDIS case.
Although the central values indicate some trend towards values smaller in magnitude as $|\vprp{b}|$ increases,
the somewhat large uncertainties and fluctuations at larger $|\vprp{b}|$ prevent us from drawing any strong conclusions.
In the range of $|\vprp{b}|\approx 0\ldots 0.4 \units{fm}$, we find a clearly non-zero negative Boer-Mulders shift
of $\langle \vect{k}_y \rangle_{u-d}^{\text{BM,SIDIS}}\approx -0.17\ldots-0.1 \units{GeV}$, for $\hat \zeta = 0.39$ and the given pion mass.
The $\hat\zeta$-dependence for $m_\pi = 518 \units{MeV}$ and $|\vprp{b}|=0.36\units{fm}$ is shown in Fig.~\ref{fig-BoerMuld_evolution}.
As for the Sivers shift, it is interesting to note that the contribution from the $\tilde A$ amplitude,
in this case $\tilde A_4 $, given by the open squares, is mostly
compatible with zero within errors, while the main signal is coming from $-R({\hat \zeta}^2) \widetilde{B}_3=-\eta\,(v\cdot P)\,R(\hat \zeta^2)\tilde b_3$,
cf.~Eqs.~(\ref{eq-ABamps}).
Finally, a comparison of the results and their $\hat\zeta$-dependences for the three different lattice ensembles is provided in Fig.~\ref{fig-BoerMuld_evolution_combined}, for a fixed $|\vprp{b}|=0.36\units{fm}$.
We find that most of the data points for the two pion masses and the two volumes are well compatible within uncertainties,
with central values of $\langle \vect{k}_y \rangle_{u-d}^{\text{BM,SIDIS}}\approx -0.2\ldots-0.1 \units{GeV}$.
While the central values show little dependence on $\hat\zeta$, the errors have to be significantly reduced before
any extrapolations towards a large Collins-Soper parameter may be attempted.
In summary, for the given ranges of parameters, our results indicate that the Boer-Mulders function is sizeable and negative for $u-d$ quarks.
Our data for the individual $u$- and $d$-quark contributions (not shown) furthermore indicate that $h^{\perp,u}_1<0$ and $h^{\perp,d}_1<0$.
Interestingly, these preliminary results are well compatible with a recent phenomenological study of the Boer-Mulders effect in SIDIS \cite{Barone:2009hw}, as well as an earlier lattice QCD study of tensor generalized parton distributions \cite{Gockeler:2006zu} in combination
with the chromodynamic lensing mechanism \cite{Burkardt:2005hp}.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-BoerMuld_lsqr-9_zetasqrlat4}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_BoerMulders_lsqr-9_zetasqrlat4}
}\hfill%
\subfloat[][]{%
\label{fig-BoerMuld_bdepend
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_BoerMuldRat_zetasqrlat4_bdepend}
}\\%
\subfloat[][]{%
\label{fig-BoerMuld_evolution}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_BoerMuldersRat_lsqr-9_evolution}
}\hfill%
\subfloat[][]{%
\label{fig-BoerMuld_evolution_combined
\includegraphics[scale=0.9]{plots/UminusD_BoerMuldRat_lsqr-9_evolution_combined_v2}
\caption[SIDIS diagram]{%
Generalized Boer-Mulders shift.\par
\subref{fig-BoerMuld_lsqr-9_zetasqrlat4}
$\eta|v|$-dependence at $m_\pi = 518 \units{MeV}$ for $\hat \zeta = 0.39$, $|\vprp{b}|=3a=0.36\units{fm}$. \\
\phantom{\subref{fig-BoerMuld_lsqr-9_zetasqrlat4}} Asymptotic results corresponding to SIDIS and DY have been extracted as in Fig. \ref{fig-Sivers_etadepend}.\par
\subref{fig-BoerMuld_bdepend}
$|\vprp{b}|$-dependence of the SIDIS results at $m_\pi = 518 \units{MeV}$, $\hat \zeta = 0.39$. \par
\subref{fig-BoerMuld_evolution}
$\hat \zeta$-dependence of the SIDIS results at $m_\pi = 518 \units{MeV}$, $|\vprp{b}|=0.36\units{fm}$. \\
\phantom{\subref{fig-BoerMuld_bdepend}} Empty squares correspond to the ratio with $\widetilde{A}_{4}$ in the numerator only. \par
\subref{fig-BoerMuld_evolution_combined}
Comparison of the $\hat \zeta$-dependence of the SIDIS results obtained from the three different lattice ensembles listed in Table \ref{tab-gaugeconfs}. \par
All error bars show statistical uncertainties only.
\label{fig-BoerMuld}%
}
\end{figure}
\subsubsection{$T$-even TMDs: The transversity $h_1$}
In the previous sections, we have discussed the T-odd Sivers and Boer-Mulders distributions, in particular their emergence
in the transition from straight to staple-shaped gauge links, i.e., as $\eta|v|$ changes from zero to large positive or negative values.
A natural question to ask is, what is the influence of final state interactions, which we
mimick on the lattice with the staple-shaped links, and which are essential for the appearance of T-odd distributions, on the T-even TMDs?
More specifically, we would like to see whether and how the T-even distributions, which are generically non-vanishing already for straight
gauge links, change during the transition to finite staple extents.
This is also of considerable interest with respect to the much less involved lattice studies of (T-even) TMDs using straight gauge links
that we have presented in \cite{Musch:2010ka}.
As we will show, there is only little difference in the transition to staple-shaped links, such that
our previous results might be of greater phenomenological
importance than initially expected for the straight ``process-independent" gauge link structures.
A suitable observable for investigating these questions is the ``generalized tensor charge" given by the ratio of the (lowest $x$-moments of the)
transversity to the unpolarized distribution, $\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)}$, defined in Eq.~(\ref{eq-h1overf1}).
The $\eta|v|$-dependence of this transversity ratio is displayed in Figs.~\ref{fig-h1Rat_lsqr-1_zetasqrlat4} to \ref{fig-h1Rat_lsqr-9_zetasqrlat4},
for different $|\vprp{b}|$ of $0.12$, $0.24$, and $0.36\units{fm}$, a pion mass of $m_\pi = 518 \units{MeV}$, and $\hat\zeta=0.39$.
We find it quite remarkable to see that $\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)}$ stays nearly constant over the full range of accessible $|\eta||v|$ in Figs.~\ref{fig-h1Rat_lsqr-1_zetasqrlat4} and \ref{fig-h1Rat_lsqr-4_zetasqrlat4}, within comparatively small statistical errors.
For $|\vprp{b}|=0.36\units{fm}$, we see little dependence apart from larger values of $|\eta||v|$ where
the signal-to-noise ratio quickly decreases. In all cases, we find indications for plateaus from $|\eta||v|\sim 3a \ldots 8a$.
As in the previous sections, we choose to average over the data in the plateau regions (solid lines), and
obtain estimates for $|\eta||v|\rightarrow\pm\infty$ from the mean of the DY and SIDIS averages by imposing the symmetry condition in $\eta |v|$.
The corresponding results are illustrated by the open diamonds.
In all considered cases, differences between $|\eta||v|\rightarrow\pm\infty$ and $|\eta||v|=0$ are barely visible within uncertainties.
In other words, lattice data for simple straight gauge links provide already a very good estimate
for the phenomenologically interesting case of infinite staple extents,
at least in the covered ranges of $\hat\zeta$ and not too large $|\vprp{b}|$.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-h1Rat_lsqr-1_zetasqrlat4}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_h1rat_lsqr-1_zetasqrlat4_FitRange3_8_v3}
}\hfill%
\subfloat[][]{%
\label{fig-h1Rat_lsqr-4_zetasqrlat4
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_h1rat_lsqr-4_zetasqrlat4_FitRange3_8_v3}
}\\%
\subfloat[][]{%
\label{fig-h1Rat_lsqr-9_zetasqrlat4
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_h1rat_lsqr-9_zetasqrlat4_FitRange3_8_v3}
}\hfill%
\subfloat[][]{%
\label{fig-h1Rat_bdepend
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_h1Rat_zetasqrlat4_bdepend_v3}
}\\%
\caption[SIDIS diagram]{%
\subref{fig-h1Rat_lsqr-1_zetasqrlat4}-\subref{fig-h1Rat_lsqr-9_zetasqrlat4} The dependence of
the transversity ratio $\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)}$, Eq.~(\ref{eq-h1overf1}), on the staple extent $\eta |v|$,
obtained at $m_\pi = 518 \units{MeV}$, $\hat \zeta = 0.39$ for three different quark separations
$|\vprp{b}|=1a=0.12\units{fm}$, $2a=0.24\units{fm}$ and $3a=0.36\units{fm}$.
Asymptotic results corresponding to SIDIS and DY have been extracted as in Fig. \ref{fig-Sivers_etadepend}, except that
we assume an even behavior of $h_{1}$ to obtain the data points plotted as empty symbols at $\eta |v| \rightarrow \pm \infty$.
The averages (lines) are obtained from the data points with staple extents in the ranges
$\eta |v| = 3a .. 8a$ and $\eta |v| = -8a .. -3a$, respectively.
Figure \subref{fig-h1Rat_lsqr-1_zetasqrlat4} might be affected by significant lattice cutoff effects due to the small quark separation
$|\vprp{b}|=a$. \\
\subref{fig-h1Rat_bdepend} The transversity ratio $\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)}$ as a function of the quark separation $|\vprp{b}|$
from the SIDIS results extracted
on the lattice with $m_\pi = 518 \units{MeV}$ for $\hat \zeta = 0.39$.
The data points lying in the shaded area below $|\vprp{b}| \approx 0.25 \units{fm}$ might be affected by lattice cutoff effects.
Error bars show statistical uncertainties only.
}
\end{figure}
The $|\vprp{b}|$-dependence of our estimates for the transversity ratio at $|\eta||v|=\pm\infty$ is displayed in Fig.\ref{fig-h1Rat_bdepend},
for $m_\pi = 518 \units{MeV}$ and $\hat\zeta=0.39$.
We observe a small, approximately linear rise of in total about $20\%$ as $|\vprp{b}|$ increases from $0.12 \units{fm}$ to about $0.6 \units{fm}$.
This is in agreement with our previous observation of a flatter $|\vprp{b}|$-dependence of the amplitude $\tilde A^{u-d}_{9m}$
compared to $\tilde A^{u-d}_{2}$ in Ref.~\cite{Musch:2010ka} on the basis of straight gauge links\footnote{Note again that $b$ in the present work corresponds to $-l$ in \cite{Musch:2010ka}.}.
Remarkably, a naive linear extrapolation of the data to $|\vprp{b}|=0$ would give a value for the tensor charge,
$g_T^{u-d}=\int\! dx\, d^2\vprp{k}\, h_1(x,\vprp{k}^2)=\tilde h_{1}^{[1](0)}(\vprp{b}\!=\!0)$, of $g_T^{u-d}\approx1.1$, which is
in very good agreement with the direct lattice calculation of this quantity using a renormalized local operator that has
been presented in \cite{Edwards:2006qx} for the same ensemble, for a scale of $\mu^2=4 \units{GeV}^2$
in the $\overline{\text{MS}}$ scheme\footnote{Note that the tensor charge is denoted by $\langle 1\rangle_{\delta q}$ in \cite{Edwards:2006qx}.}.
Figure \ref{fig-h1Rat_evolution} shows the transversity ratio as a function of the Collins-Soper parameter, for the same pion mass as before but a fixed $|\vprp{b}|=0.36\units{fm}$.
The $\hat\zeta$-dependence turns out to be rather flat over the full range of accessible values.
It is interesting to note that, in contrast to the T-odd distributions discussed before,
the amplitude $\tilde A_{9m}$ (open circles) provides $\sim100\%$ of the total results, while
the contribution from $R({\hat \zeta}^2) \widetilde{B}_{15}$, Eq.~(\ref{eq-ABamps}), as well as from $\widetilde{B}_{17}$ through Eq.~(\ref{eq-A9m}),
is negligible within errors, over the full range of $\hat\zeta$.
Finally, we show a comparison of our results for $(\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)})^{u-d}$ obtained for the different ensembles in Fig.~\ref{fig-h1Rat_evolution_combined}.
As before, the data points for the two values of the pion mass and the different volumes agree within uncertainties.
On the basis of the comparatively good signal-to-noise ratio for this observable, we conclude that the $\hat\zeta$-dependence
is in this case rather flat and very well compatible with a constant behavior,
$(\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)})^{u-d}(\hat\zeta)\approx 1.2\pm0.1$, at least for $\hat\zeta\le0.8$ and the given parameters.
It would be interesting to investigate in future lattice studies whether this
constant behavior persists as one approaches larger Collins-Soper parameters.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-h1Rat_evolution}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_h1Rat_lsqr-9_evolution_v3}
}\hfill%
\subfloat[][]{%
\label{fig-h1Rat_evolution_combined
\includegraphics[scale=0.9]{plots/UminusD_h1Rat_lsqr-9_evolution_combined_v3}
}\\%
\caption[SIDIS diagram]{%
Evolution with respect to $\hat \zeta$ for the transversity ratio $\tilde h_{1}^{[1](0)}/\tilde f_1^{[1](0)}$ at a quark separation of $|\vprp{b}|=3a=0.36\units{fm}$.
Figure \ref{fig-h1Rat_evolution} shows the SIDIS results obtained at $m_\pi = 518\units{MeV}$. The solid data points correspond to the full result,
and empty symbols to the result obtained with just $\widetilde{A}_{9m}$ in the numerator.
Figure \ref{fig-h1Rat_evolution_combined} displays the full results for all ensembles listed in Table \ref{tab-gaugeconfs}.
\label{fig-h1Rat_aevolution}%
}
\end{figure}
\subsubsection{$T$-even TMDs: The generalized worm gear shift from $g_{1T}$}
As a final example, we study in this section the generalized shift defined
in Eq.~(\ref{eq-geng1TShift}), which is essentially given by the T-even
TMD $g_{1T}$. Figure~\ref{fig-g1TRat_etadepend} shows
$(\tilde g_{1T}^{[1](1)}/\tilde f_1^{[1](0)})^{u-d}$ as a function of
$\eta |v|$ for $m_\pi = 518 \units{MeV}$, $\hat \zeta = 0.39$
and two values of $|\vprp{b}|$.
Within uncertainties, we observe, as expected, an approximate symmetry with respect to the sign of $\eta |v|$.
Furthermore, we find overall only little dependence on the staple extent for $|\vprp{b}|=0.12\units{fm}$.
At larger $|\vprp{b}|$, the signal-to-noise ratio quickly decreases as $|\eta| |v|$ becomes larger.
Still, we find indications that the results stabilize in the region $|\eta| |v| = 3a \ldots 8a$, which we choose as our plateau
region for the computation of average values, as discussed in the previous section.
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-g1TRat_lsqr-1_zetasqrlat4}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_g1Trat_lsqr-1_zetasqrlat4_v3}
}\hfill%
\subfloat[][]{%
\label{fig-g1TRat_lsqr-9_zetasqrlat4
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_g1Trat_lsqr-9_zetasqrlat4_v3}
}\\%
\caption[SIDIS diagram]{%
Dependence of the generalized $g_{1T}$ shift on the staple extent $\eta |v|$,
obtained at $m_\pi = 518 \units{MeV}$, $\hat \zeta = 0.39$ for two different quark separations
$|\vprp{b}|=1a=0.12\units{fm}$ and $|\vprp{b}|=3a=0.36\units{fm}$.
Asymptotic results corresponding to SIDIS and DY have been extracted as in Figs. \ref{fig-h1Rat_lsqr-1_zetasqrlat4} to \ref{fig-h1Rat_lsqr-9_zetasqrlat4}.
Error bars show statistical uncertainties only.
Figure \subref{fig-g1TRat_lsqr-1_zetasqrlat4} might be affected by significant lattice cutoff effects due to the small quark separation $|\vprp{b}|=a$.
\label{fig-g1TRat_etadepend}%
}
\end{figure}
As before, the averages serve as approximations for the asymptotic results at $\eta |v| = \pm\infty$, i.e., corresponding to infinite staple extents.
The dependence of these asymptotic values on $|\vprp{b}|$ is displayed in Fig.~\ref{fig-g1TRat_bdepend}.
Although a small curvature in the central values can be observed, the results are overall rather stable within errors, with
$\langle \vect{k}_x \rangle_{u-d}^{{g_{1T}}}=(\tilde g_{1T}^{[1](1)}/\tilde f_1^{[1](0)})^{u-d}\approx0.16\ldots 0.21 \units{GeV}$.
In Fig.~\ref{fig-g1TRat_evolution}, we show the dependence of the generalized $g_{1T}$ shift on the Collins-Soper parameter,
for a pion mass of $m_\pi = 518 \units{MeV}$ and a fixed $|\vprp{b}|=0.36\units{fm}$.
As $\hat\zeta$ increases, one observes a slight trend towards values that are smaller in magnitude, although
it is difficult to draw any strong conclusions in view of the present uncertainties.
Similar to the case of the transversity ratio of the previous section, we find that
essentially the full signal is due to the amplitude $\tilde A_7$, while the contribution from
$R({\hat \zeta}^2) \widetilde{B}_{13}$, cf. Eq.~(\ref{eq-ABamps}), is compatible with zero within errors.
Finally, Fig.~\ref{fig-g1TRat_evolution_combined} gives an overview of our results as functions of $\hat\zeta$, obtained
for the three considered ensembles, for $|\vprp{b}|=0.36\units{fm}$.
Apart from $\hat\zeta=0$, the data points for the two pion masses and volumes clearly overlap within uncertainties.
Taking into consideration the results for $m_\pi = 369 \units{MeV}$ and a spatial volume of $28^3$ (given by the filled diamonds),
the data are overall compatible with a constant behavior, although more statistics is necessary to establish a clear trend in $\hat\zeta$.
Altogether, we observe a sizeable positive generalized transverse shift in the range of
$\langle \vect{k}_x \rangle_{u-d}^{{g_{1T}}}=(\tilde g_{1T}^{[1](1)}/\tilde f_1^{[1](0)})^{u-d}\approx0.15\ldots\ 0.25 \units{GeV}$,
for $\hat\zeta=0\ldots0.8$ and the given parameters.
We note that these values are in good agreement with our previous analyses on the basis of straight gauge links \cite{Hagler:2009mb,Musch:2010ka}.
\begin{figure}[btp]
\centering%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_g1TRat_zetasqrlat4_bdepend_v3}
\caption{%
Generalized $g_{1T}$ shift for $|\eta| |v| = \infty$ as a function of the quark separation $|\vprp{b}|$
from the SIDIS and DY results extracted
on the lattice with $m_\pi = 518 \units{MeV}$ for $\hat \zeta = 0.39$.
The data points lying in the shaded area below $|\vprp{b}| \approx 0.25 \units{fm}$ might be affected by lattice cutoff effects.
Error bars show statistical uncertainties only.
\label{fig-g1TRat_bdepend}%
}
\end{figure}
\begin{figure}[btp]
\centering%
\subfloat[][]{%
\label{fig-g1TRat_evolution}%
\includegraphics[scale=0.9]{plots/m020_run38_UminusD_g1TRat_lsqr-9_evolution_v3}
}\hfill%
\subfloat[][]{%
\label{fig-g1TRat_evolution_combined
\includegraphics[scale=0.9]{plots/UminusD_g1TRat_lsqr-9_evolution_combined_v3}
}\\%
\caption[SIDIS diagram]{%
Evolution with respect to $\hat \zeta$ for the generalized $g_{1T}$ shift at a quark separation of $|\vprp{b}|=3a=0.36\units{fm}$.
Figure \ref{fig-g1TRat_evolution} shows the results obtained at $m_\pi = 518\units{MeV}$ for the SIDIS and DY limit $|\eta|| v|\rightarrow \infty$. The solid data points correspond to the full result
and empty symbols to the result obtained with just $\widetilde{A}_{7}$ in the numerator.
Figure \ref{fig-g1TRat_evolution_combined} displays the full results for all ensembles listed in Table \ref{tab-gaugeconfs}.
\label{fig-g1TRat_aevolution}%
}
\end{figure}
\section{Summary and Conclusions}
\label{sec-summary}
We have presented an exploratory study of quark transverse momentum
distributions in the nucleon in full lattice QCD employing non-local operators
with staple-shaped gauge links (Wilson lines).
Compared to our earlier works \cite{Hagler:2009mb,Musch:2010ka}, the use of staple-shaped instead of straight link paths allowed us for the first time to systematically access the naively time-reversal odd (T-odd) observables, in particular the amplitudes related to the Sivers and the Boer-Mulders TMDs.
In the framework of QCD factorization theorems, the path dependence
corresponds to a process dependence that leads to the famous sign difference
between the T-odd TMDs for the SIDIS and the DY processes. In our study, we
were able to distinguish the SIDIS and DY cases through the relative
orientation of the nucleon momentum $P$ and the vector $\eta v$ that
characterizes the direction and extent of the staple on the lattice,
cf. Fig.~\ref{fig-link-staple}. It is important to keep in mind that TMDs
defined with non-light-like staple vectors $v$, as required on the lattice,
will additionally depend on the Collins-Soper evolution parameter, here denoted by $\hat\zeta$.
In order to avoid additional soft factors in the formal definition of the TMDs, we have concentrated
on the Sivers, Boer-Mulders, and worm-gear ($g_{1T}$) generalized transverse momentum shifts and the generalized tensor charge.
Since the generalized shifts and tensor charge are defined in terms of ratios of TMDs, potential soft factors
as well as the renormalization constants cancel out.
Our numerical results, obtained for three different ensembles with pion masses
$m_\pi = 369 \units{MeV}$ and $m_\pi = 518 \units{MeV}$, as well as spatial
lattice volumes of $20^3$ and $28^3$, are very promising: We find clearly
non-zero, sizeable signals for all observables we considered. The expected
anti-symmetry (change of sign) for T-odd quantities in $\eta |v|$ is
fulfilled within statistical uncertainties. In contrast, for the T-even
quantities we observe little systematic dependence on the staple direction
and extent. As the staple extents are increased, our data appear to approach
plateaus. Averages of the plateau values then provide estimates for the limit
of infinite staple extents, $\eta |v|\rightarrow\pm\infty$, which is formally
required for all phenomenologically relevant TMDs.
The physical length scale beyond which the influence of the gauge link
extent diminishes is of the order of $0.4\units{fm}$ for all cases considered. This observation
invites speculations as to the physical background of this scale, e.g., an interpretation as color correlation length.
The scale might also be related to a mass gap in the spectrum.
If $v$ is interpreted as the Euclidean time direction, the legs of the staple-shaped gauge link resemble static quark propagators.
Considering our three-point function in this rotated frame of reference suggests that the plateau region is reached when the propagation time $|\eta v|$ of the static quark pair is large enough to suppress contributions from excited states sufficiently.
Our numerical extrapolations to infinite staple extents,
$\eta |v|\rightarrow\pm\infty$, represent first predictions for the signs and
approximate sizes of the generalized transverse shifts from lattice QCD.
In particular, we find strong indications that the T-odd Sivers and
Boer-Mulders TMDs are both sizeable and negative for the isovector,
$u-d$ quark combination in the case of SIDIS.
Within statistical errors, we do not observe any clear trend in the data for
the transverse shifts as functions of the Collins-Soper evolution parameter
$\hat\zeta$ in the range $\hat\zeta\sim0 \ldots 0.8$.
For the T-even generalized tensor charge, which shows a much better signal-to-noise ratio and less scatter of the data points,
we can tentatively conclude that it is approximately constant in $\hat\zeta$ for the accessible parameter ranges.
We stress, however, that more quantitative predictions with respect to phenomenological analyses
of SIDIS and DY experiments on the basis of QCD factorization will require much larger Collins-Soper parameters $\hat\zeta \gg 1$.
For the TMD ratios discussed in this study, large $\hat\zeta$ can in principle be accessed through larger nucleon momenta.
In practice, this represents a considerable challenge due to quickly decreasing signal-to-noise ratios and potentially
significant finite volume effects at higher $P$.
Still, we expect that future lattice results for an extended range of momenta
and with improved statistics will be very useful to establish trends in
$\hat \zeta$, eventually allowing extrapolation into the region where
factorization theorems and related evolution equations are applicable.
\begin{acknowledgments}
We thank Harut Avakian, Gunnar Bali, Alexei Bazavov, Vladimir Braun, Markus Diehl, Robert Edwards, Meinulf G\"ockeler, Barbara Pasquini, Alexei Prokudin, David Richards and Dru Renner for helpful discussions and suggestions.
We are grateful to the LHP collaboration for providing their lattice quark
propagators to us, and for technical advice, as well as to the MILC
collaboration for use of their Asqtad configurations.
Our calculations, which relied on the Chroma software suite
\cite{Edwards:2004sx}, employed computing resources provided by
the U.S.~Department of Energy through USQCD at Jefferson Lab.
The authors acknowledge support by the Heisenberg-Fellowship program of the DFG
(Ph.H.), SFB/TRR-55 (A.S.) and the U.S.~Department of Energy under grants
DE-FG02-96ER40965 (M.E.) and DE-FG02-94ER40818 (J.N.). M.E.~furthermore is
grateful to the Jefferson Lab Theory Center for its generous support and
hospitality during Fall 2011, which proved invaluable for the progress of
this project. Authored by Jefferson Science Associates, LLC under
U.S. DOE Contract No. DE-AC05-06OR23177.
The U.S. Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce this manuscript for U.S. Government purposes.
\end{acknowledgments}
|
\section{Introduction}
Solid state realization of q-bits that do not decohere easily is a challenging
task in the field of Quantum Computation. Topological defects in strongly
correlated quantum many body systems are protected from decoherence and have
been suggested as q-bit candidates \cite{qcgen1,qcgen2,qcgen3,qcgen4,qcgen5}.
In this context, Kitaev constructed a remarkable two dimensional quantum spin
model that exhibits abelian and non-abelian anyons and is exactly solvable for
its spectrum \cite{kit1,kit2}. It was later shown that the spin correlation
functions are also exactly solvable \cite{short}. This model is also extremely
interesting form point of view of frustrated spin models and the physics of the
resonating valence bond (RVB)
states\cite{rvb-dm1,rvb-dm2,rvb-dm3,rvb-dm4,rvb-dm5}. It realizes, in an exact
fashion, the phenomenon of quantum number fractionization and emergent gauge
fields that were conjectured and approximately realized in 2D models for RVB
states or quantum spin liquids\cite{frac1,frac2}. It has also been shown that
the Jordan-Wigner transformation in this model yields a local fermionic
theory\cite{jw1,jw2,jw3,jw4}. This makes the Kitaev model an important one
that warrants further investigation; and no wonder that an extensive body of
research \cite{othkit1,othkit2,othkit3,othkit4,othkit5} has already been
carried out exploring its many fascinating aspects. Kitaev showed that the
model has a natural formulation in terms of a Majorana fermion interacting with
Z$_2$ gauge fields. The remarkable feature of the Kitaev model is that the
gauge fields turn out to be static. This greatly simplifies the dynamics
leading to the exact computation of the spectrum and spin-spin correlation
functions.
The gauge theory of spin-$\frac{1}{2}$ models has a recent history. It was
initially formulated \cite{bza} in the context of strongly correlated
electronic systems such as a spin-$\frac{1}{2}$ Mott insulator, as a way of
implementing the single electron occupancy constraint,
\begin{equation}
\label{u1gl}
\sum_{\sigma}\left(c^\dagger_{i\sigma}c_{i\sigma}-1\right)\vert\psi\rangle=0.
\end{equation}
This equation is the Gauss law constraint for a $U(1)$ lattice gauge theory.
It resulted in a strongly interacting U(1) gauge theory formalism of
spin-$\frac{1}{2}$ models. Soon it was realized that in spin-$\frac{1}{2}$
systems, the $U(1)$ gauge invariance always implied an $SU(2)$ invariance
\cite{su21,su22,su23}. The consequent $SU(2)$ gauge theory formalism was found
to be useful in the context of relating apparently different mean field
solutions of the model. The extended Hilbert space of the spin-$\frac{1}{2}$
system is much smaller than that of normal lattice gauge theory where the gauge
degrees of freedom on the links are $SU(2)$ group elements. Consequently, it
was shown that the essential physics of the spin-$\frac{1}{2}$ system is
captured by a Z$_2$ gauge theory \cite{bt,martson}, where the Z$_2$ gauge group is the
center of the original $SU(2)$ gauge group. The Z$_2$ gauge theory formalism
has been effectively used to bring out the physics of quantum number
fractionization in spin-$\frac{1}{2}$ systems \cite{frac2}.
In this paper we follow the route charted out above in the context of the
Kitaev model and show that the Z$_2$ gauge theory can indeed be thought of as
the center of the $SU(2)$ gauge theory of RVB theory. This sheds light on
Kitaev's assertion \cite{kit1} that the model represents the same universality
class of topological order as RVB. We examine the degeneracy of states in the
system defined with periodic boundary conditions in detail. We show that
this degeneracy, which characterises the topological order in the system,
arises from the so called large gauge transformations. Namely, gauge field
configurations which correspond to the same flux configuration that are not
related to each other by local gauge transformations. These topologically
distinct gauge field configurations can be labelled by the value of the Wilson
loops that wind around the torus in the two different directions. We
find these gauge configurations for the ground state and demonstrate the
four-fold degeneracy of the ground state by explicity computing the energies.
Further, we generalise this proof to all eigenstates by constructing the
operators that generate the large gauge transformations and showing that
they do not change the energy of the system in the thermodynamic limit.
The rest of the paper is organised as folows. Section (\ref{km}) reviews the
Kitaev honeycomb model with its main features and some mathematical notions
which are used in later sections. The $SU(2)$ gauge symmetry of the model is
reviewed in section (\ref{su22z2}) where we show that Kitaev's choice of the
representation of the spin operators in terms of Majorana fermions amounts to a
$SU(2)$ gauge fixing procedure which fixes the gauge upto the center Z$_2$
gauge transformations. In section (\ref{z2gtkm}), we construct a generalized
Jordan-Wigner transformation and show that it is a Z$_2$ gauge fixing
transformation in the Kitaev formalism. We derive the fermionic Hamiltonian and
simplify the conserved quantities in terms of gauge invariant Jordan-Wigner
operators. In section (\ref{gsd}), we explicitly work out the four fold
degeneracies of the Kitaev model on a torus. We derive the fermionic spectrum
for flux free sector in (\ref{gsd1}) and show that it leads to four degenerate
ground states in thermodynamic limit. Following this in section (\ref{tpg}), we
demonstrate that every eigenstate of the Kitaev model has four fold degeneracy.
To this end we derive four mutually anticommuting operators that commute with
the hamiltonian in the thermodynamic limit. The minimum dimension of the
representation of the 4-dimensional Clifford algebra is 4. Thus we are able to
prove that all eigenstates are at least four fold degenerate. We summarize our
results in the section (\ref{dis}).\\
\section{The Kitaev model}
\label{km}
\subsection{The Hamiltonian}
\label{modeldef}
The Kitaev model is a spin-$\frac{1}{2}$ system on a honeycomb
lattice. The Hamiltonian is
\begin{equation}
\label{1}
H=-J_{x}\sum_{{\langle ij\rangle}_x}\sigma^{x}_{i}\sigma^{x}_{j}
-J_{y}\sum_{{\langle ij\rangle}_y}\sigma^{y}_{i}\sigma^{y}_{j}
-J_{z}\sum_{{\langle ij\rangle}_z}\sigma^{z}_{i}\sigma^{z}_{j},
\end{equation}
where $i,j$ run over the sites of the honeycomb lattice,
$\langle ij\rangle_a,~a=x,y,z$ denotes the nearest
neighbor links oriented in the $a$'th direction as shown in
Fig. \ref{kmd}. We will be working with periodic boundary conditions which
are defined as follows. The honeycomb lattice is a triangular lattice
with a basis of two sites. The sites of the triangular lattice are given by,
\begin{equation}
\label{tlsites}
{\bf R}_{m,n}=m{\bf e}_1+n{\bf e}_2,
\end{equation}
where $m,~n$ are integers and ${\bf e}_1\cdot{\bf e}_2=-\frac{1}{2},~
{\bf e}_1\cdot{\bf e}_1=1={\bf e}_2\cdot{\bf e}_2$. The label $i$ of the sites
of the honeycomb lattice therefore stands for $(m,n,\alpha)$ where $\alpha=a,b$ is
the sub lattice label. The periodic boundary conditions are then defined by,
\begin{equation}
\sigma^a_{m,n,\alpha}=\sigma^a_{m+M,n+N,\alpha}.
\end{equation}
\begin{figure}[h!]
\center{\hbox{\epsfig{figure=hexagon4.eps,height=1.2in,width=2.5in}}}
\caption{Honeycomb lattice and the Kitaev model}
\label{kmd}
\end{figure}
\subsection{The conserved quantities}
\label{spconsq}
There is a conserved quantity associated with every plaquette of
the lattice. If the plaquette is denoted as $p$ and its vertices
labelled as shown in Fig. \ref{kmd}, then, following Kitaev's notation,
the conserved quantity is,
\begin{equation}
\label{bpdef}
B_p=\sigma_1^x\sigma_2^z\sigma_3^y\sigma_4^x\sigma_5^z\sigma_6^y.
\end{equation}
We have $B_p^2=1$ implying that the $B_p$s can take values $\pm 1$.
It is clear that any product of the $B_p$s will also commute with the
Hamiltonian. In fact there is a conserved quantity associated with every
closed self avoiding loop, $C$, on the lattice defined the following way.
At every site, the path will pass through two of the
three bonds that emanate from it. We call these two bonds as the tangential
bonds and the third one the normal bond. We associate two tangential vectors
at each site, ${\bf{\hat t}}_{1i}$ and ${\bf{\hat t}}_{2i}$
which are either ${\bf{\hat x}}$, ${\bf{\hat y}}$ or ${\bf{\hat z}}$
according to the direction of the incoming bond and the outgoing bond
respectively. We then define a normal vector ${\bf{\hat n}}_i$ as
\begin{equation}
\label{nidef}
{\bf{\hat n}}_i\equiv{\bf{\hat t}}_{1i}\times{\bf{\hat t}}_{2i}.
\end{equation}
If the sites of $C$ are $i_1,i_2,.....i_N$, then the
conserved quantity associated with it is,
\begin{equation}
\label{bcdef}
B_C=\prod_{n=1}^N\left({\bf{\hat n}}_{i_n}\cdot{\bf\sigma}_{i_n}\right).
\end{equation}
It can be checked that,
\begin{equation}
[B_C,H]=0,~~~~B_C^2=1.
\end{equation}
We will call $C$ topologically trivial if it can be written as a product
of $B_p$s. On the torus, we have two loops which wind the torus around
in the two directions which cannot be expressed as a product of $B_p$s.
One cannot be obtained from the other by multiplication by $B_p$s. We will
call these two (Wilson)loops $W_1$ and $W_2$.
All the $B_p$s are not independent due to the identity,
\begin{equation}
\label{bpid}
\prod_p B_p=1.
\end{equation}
Thus there are $N_p-1$ independent $B_p$s, where $N_p=MN$ is the number
of plaquettes. Together with $W_1$ and $W_2$, we have a total
of $N_p+1$ conserved quantities on the torus. These two loop operators
account for the four fold degeneracy on a torus.
\section{From $SU(2)$ to Z$_2$}
\label{su22z2}
\subsection{$U(1)$ and $SU(2)$ gauge symmetry}
Interacting quantum spin systems often lead to spontaneously broken symmetric
states such as a ferro, antiferro or spiral magnetic states. Low energy physics of these ordered states are captured by the well known spin wave approximations, reasulting in Goldstone mode type bosonic low energy effective theories \cite{fradkin}. A quantum spin liquid, on the other hand, has no classical long range order. Experiments in LaCuO$_{4}$ and other low spin Mott insulators, according to Anderson, indicated possible presence of neutral fermionic excitations in a quantum spin liquid \cite{nspin1,nspin2,nspin3}. A theory to describe such a quantum spin liquid or resonating valence bond (RVB) state needed a paradigm shift from spin wave theory. It was also clear that a quantum spin liquid, in view of different possible phase coherence among disordered spin configurations, could offer a variety of quantum spin liquid states to be realized in nature. RVB gauge theory attempted to capture these new possibilties, through an approach involving enlarged Hilbert space and emergent gauge fields in strongly correlated electron systems. Through the work of Wen\cite{WenPSG} and others it has become clear that there is a plethora of spin liquid phases, characterized by \textit{quantum order and projective symmetry groups.}
The spin-$\frac{1}{2}$ Hilbert space can be realized as the subspace
of the Hilbert space of two fermions defined by the constraint,
\begin{equation}
\label{u1cons}
\left(\sum_{\sigma=\uparrow\downarrow}~c^\dagger_\sigma c_\sigma -1\right)
\vert\psi\rangle=0,
\end{equation}
where $c^\dagger_\sigma$ and $c_\sigma$ are the fermion creation and
annihilation operators. As mentioned earlier, Eq.~(\ref{u1cons}) can
be looked upon as the Gauss law constraint for a $U(1)$ gauge theory. The
LHS of the equation being the generator of the following $U(1)$ gauge
transformations on the fermion operators,
\begin{equation}
\label{u1trans}
c_\sigma\rightarrow e^{i\Omega}c_\sigma,~~~
c^\dagger_\sigma\rightarrow e^{-i\Omega}c^\dagger_\sigma.
\end{equation}
The spin operators,
\begin{equation}
\label{spindef}
S^a=\frac{1}{2}
c^\dagger_\sigma\sigma^a_{\sigma\sigma^\prime}c_{\sigma^\prime},
\end{equation}
are then the gauge invariant observables of the theory. $\sigma^a$ are the
Pauli spin matrices.
The single occupancy constraint in the spin-$\frac{1}{2}$ theory implies
that a spin-$\uparrow$ hole is the same as a spin-$\downarrow$ particle in
the physical space. This can be mathematically expressed as an $SU(2)$
gauge invariance. It is convenient to express this symmetry in terms of
a matrix of the fermion operators
\begin{equation}
\label{smatdef}
{ \Psi}\equiv\left(\begin{array}{cc}
c_\uparrow& -c^\dagger_\downarrow\\
c_\downarrow& c^\dagger_\uparrow
\end{array}\right).
\end{equation}
In terms of this matrix, the spin operators are given by,
\begin{equation}
\label{gtgs}
S^a=\frac{1}{4}{\rm tr}~{ \Psi}^\dagger \sigma^a { \Psi}.
\end{equation}
The generators of the $SU(2)$ gauge transformation are given by,
\begin{equation}
\label{sps}
\tilde S^a=-\frac{1}{4}{\rm tr}~{ \Psi}\sigma^a{ \Psi}^\dagger.
\end{equation}
The ${ \Psi}$ matrix transforms under the $SU(2)$ spin and $SU(2)$
gauge transformations as,
\begin{equation}
\label{smtrans}
{ \Psi}\rightarrow U_S{ \Psi}U^\dagger_G.
\end{equation}
Where $U_S$ and $U_G$ are $SU(2)$ matrices representing the spin and
gauge transformations respectively. It is clear from equations~
(\ref{gtgs}), (\ref{sps}) and (\ref{smtrans}) that the spin operators are
gauge invariant and the generators of gauge transformations are spin singlet.
The constraint in Eq.~(\ref{u1cons}) is exactly equivalent to the $SU(2)$
Gauss law,
\begin{equation}
\label{su2gl}
\tilde S^a\vert\psi\rangle=0.
\end{equation}
Before we close this section, we wish to mention that the above gauge theory formalism offers a possible way to understand quantum spin liquid states as and when they exist. This formalism does not gaurantee a simple gauge theory structure at all energy scales in the physics of the problem. It only suggests that in some systems (for some Hamiltonians) at low energy scales there could be emergent gauge fields and interesting consequences of quantum number fractionization, quantum order etc.
At high energy scales gauge fields interact and it is no more simple or useful to talk in terms of emergent gauge fields. As we will see soon, the Hamiltonian invented by Kitaev on a honeycomb lattice is very special. It offers static Z$_2$ gauge fields and makes the Z$_2$ gauge theory meaningful at all energy scales.
\subsection{Majorana fermions and the Z$_2$ theory}
We can make connection to Kitaev's representation of the spins by writing,
\begin{equation}
\label{kitrep}
c_\uparrow=\frac{c_y-ic_x}{2},~~~c_\downarrow=-\frac{c-ic_z}{2}.
\end{equation}
where $c,c_x,c_y~{\rm and}~c_z$ are Majorana fermions. The single occupancy
constraint, Eq.~(\ref{u1cons}) reduces to exactly Kitaev's form,
\begin{equation}
\label{kitcons}
cc_xc_yc_z=1.
\end{equation}
Kitaev's representation of the spins then get written as,
\begin{equation}
\label{kitsprep}
\frac{i}{2}cc_a=S^a-\tilde S^a.
\end{equation}
Note that these three operators are not equal to the gauge invariant spin
operators in the extended Hilbert space but are exactly equivalent to them
in the physical Hilbert space. Substituting the expressions in Eq.~(\ref{kitsprep}) for the spin operators in the Hamiltonian is then equivalent
to adding gauge fixing terms. Since the $SU(2)$ gauge generators are invariant
under the Z$_2$ center of the gauge group, these terms only fix the gauge
upto the central Z$_2$ group represented by $e^{i2\pi\tilde S^3}$. The
Hamiltonian will therefore continue to have a Z$_2$ gauge symmetry.
The simple example of a spin-$\frac{1}{2}$ in a magnetic field illustrates
these issues. If we take the Hamiltonian to be
\begin{equation}
\label{spbham1}
H=B~S^3.
\end{equation}
then the theory has $SU(2)$ gauge symmetry, the degenerate ground states in
the extended space are
\begin{equation}
\label{su2gs}
\vert GS\rangle=\alpha c^\dagger_\downarrow\vert 0\rangle
+\beta\vert 0\rangle
+\gamma c^\dagger_\uparrow c^\dagger_\downarrow\vert 0\rangle,
\end{equation}
for arbitrary $\alpha, \beta~{\rm and}~\gamma,
~\vert\alpha\vert^2+\vert\beta\vert^2+\vert\gamma\vert^2=1$. The last two
states in the RHS of Eq.~(\ref{su2gs}) transform as a doublet under the
$SU(2)$ gauge symmetry. Gauge averaging therefore projects out the ground
state in the physical subspace, namely $c^\dagger_\downarrow\vert 0\rangle$.
If the Hamiltonian is taken to be,
\begin{equation}
\label{spbham2}
H=iB~cc_z=B~\left(S^3-\tilde S^3\right).
\end{equation}
The theory has only Z$_2$ gauge invariance, the degenerate ground state in the
extended Hilbert space are,
\begin{equation}
\label{z2gs}
\vert GS\rangle=\alpha c^\dagger_\downarrow\vert 0\rangle
+\gamma c^\dagger_\uparrow c^\dagger_\downarrow\vert 0\rangle.
\end{equation}
The second state in the above equation transforms non-trivially under the
Z$_2$ gauge transformation. Thus again, under gauge averaging, the ground
state in the physical sector is projected out.
Thus Kitaev's representation of the spin operators can be interpreted
as adding gauge fixing terms to the $SU(2)$ gauge invariant hamiltonian
which leave an unbroken (unfixed) Z$_2$ gauge symmetry.
\section{The Z$_2$ gauge theory of the Kitaev model}
\label{z2gtkm}
\subsection{The Hamiltonian}
Following Kitaev \cite{kit1}, we write the hamiltonian in terms of the
Majorana fermions,
\begin{equation}
\label{z2ham1}
\tilde H=\sum_{a=1}^3\sum_{\langle ij\rangle_a}ic_iu_{\langle ij\rangle_a}c_j,
\end{equation}
where the link variables are defined as
$u_{\langle ij\rangle_a}=ic_{ai}c_{aj}$. It
is natural to express them in terms of the bond fermions \cite{short}
defined as,
\begin{equation}
\label{bfdef}
\chi_{\langle ij\rangle_a}\equiv \frac{c_{ai}+ic_{aj}}{2}.
\end{equation}
The link variables are then given in terms of the occupancy number of the bond
fermions,
\begin{equation}
\label{bfgf}
u_{\langle ij\rangle_a}=
2~\chi^\dagger_{\langle ij \rangle_a}\chi_{\langle ij\rangle_a}-1.
\end{equation}
It is easy to see that,
\begin{eqnarray}
\label{uijprop1}
u_{\langle ij\rangle_a}^2&=&1, \nonumber\\
\label{uijprop2}
[u_{\langle ij\rangle_a},H]&=&0.
\end{eqnarray}
Thus the link variables can be interpreted as static Z$_2$ gauge fields.
It is remarkable that at one shot Kitaev hamiltonian has been solved exactly for the entire many body spectrum ! Infact, two related phenomena occur:
i) the Z$_{2}$ gauge theory is exact at all energy scales and ii) the enlarged Hilbert space gets decomposed into sectors that are identical gauge copies having the same energy eigen values (Fig. \ref{hlpt1} ). The Hilbert space enlargement does not produce any unphysical state, but only gauge copies. In the standard U(1) RVB gauge theory for a Heisenberg antiferromagnet, for example, it is easy to see how unphysical states are brought in by Hilbert space enlargement. For example, an unphysical state containing M doubly occupied and M empty sites gives spectrum of a Heisenberg antiferromagnet containing 2M missing sites.
\begin{figure}[h!]
\center{\hbox{\epsfig{figure=Hilberspace.fig.eps,height=1.7in,width=2.9in}}}
\caption{How the enlarged Hilbert space for Kitaev Model become just Gauge Copies.}
\label{hlpt1}
\end{figure}
\subsection{The Jordan-Wigner transformation and Z$_2$ gauge fixing}
A remarkable feature of the Kitaev model is that the Jordan-Wigner
transformation yields a local fermionic Hamiltonian \cite{jw1,jw2,jw3,jw4}. In this
section we show that the Jordan-Wigner transformation in the Kitaev model
can be interpreted as a Z$_2$ gauge fixing procedure resulting in gauge
invariant (gauge fixed) Majorana and bond fermions. The choice of the
Jordan Wigner path amounts to a choice of the Z$_2$ gauge.
\subsubsection{The Jordan-Wigner fermionisation}
We define the Jordan-Wigner transformation as follows ~\cite{abhinav}. Take any Hamilton
path on the lattice defined by a sequence of sites $i_n,~n=1,\dots,N_S$,
where $N_S$ is the number of sites in the lattice. The path will classify
each bond as normal or tangential as defined in section \ref{spconsq}.
The normal bonds will form a dimer covering of the lattice. For a given site
`$i$' we attach three vectors, two tangential vectors denoted by $\rm \hat{\bf{t}}_{1i},~ \hat{\bf{t}}_{2i}$ and one normal vectors $\rm \hat{\bf{n}}_{i}$, such that they follow Eq.~(\ref{nidef}). Then we can define the Jordon-Wigner transformations in the following compact way,
\subsubsection{Gauge invariant Jordan-Wigner fermions}
To define the Jordan-Wigner transformations we associate two Majorana fermions ($\eta_{i_n}$ and $\xi_{i_n}$) at a given site `$i_n$'. These Jordan-Wigner (Majorana) fermions are defined in terms of the gauge invariant spin-operators in the following way ,
\begin{eqnarray}
\label{etadef}
\eta_{i_n}&=&{\bf{\hat t}}_{1i_n}\cdot{\mathbf\sigma}_{i_n}~\mu_{i_n}. \\
\label{xidef}
\xi_{i_n}&=&{\bf{\hat t}}_{2i_n}\cdot{\mathbf\sigma}_{i_n}~\mu_{i_n}. \\
\mu_{i_n}&=&
\prod_{m=1}^{n-1}\left({\bf{\hat n}}_{i_m}\cdot{\mathbf\sigma}_{i_m}\right).
\label{mudef}
\end{eqnarray}
It can be easily checked that the above definitions refer to the usual anti-commutations relations for Majorana fermions,
\begin{eqnarray}
\{\eta_i,\xi_j\}&=&2\delta_{ij}, \nonumber \\
\{\xi_i,\eta_j\}&=&2\delta_{ij}, \nonumber \\
\{\eta_i,\xi_j\}&=&0 \label{jwacr1}.
\end{eqnarray}
Since the Jordan-Wigner fermions are constructed entirely from the spin
operators, they are manifestly gauge invariant. However, it is also
interesting to see this by rewriting equations (\ref{xidef}) and
(\ref{etadef}) in terms of the original Majorana fermions and gauge fields,
\begin{eqnarray}
\label{etadef1}
\eta_{i_n}&=&ic_{i_n}~\left(u_{i_{n-1}i_{n-2}}
\dots u_{i_2i_1}\right)~{\bf{\hat t}}_{1i_1}\cdot{\bf c}_{i_1}.\\
\label{xidef1}
\xi_{i_n}&=&i{\bf c}_{i_n}\cdot{\bf{\hat n}}_{i_n}
~\left(u_{i_{n-1}i_{n-2}}
\dots u_{i_2i_1}\right)~{\bf{\hat t}}_{1i_1}\cdot{\bf c}_{i_1}.
\end{eqnarray}
The transformation can be inverted to write the spins in terms of the fermions,
\begin{eqnarray}
{\bf{\hat n}}_{i_n}\cdot{\mathbf\sigma}_{i_n}&=&i\eta_{i_n}\xi_{i_n}.\\
{\bf{\hat t}}_{1i_n}\cdot{\mathbf\sigma}_{i_n}&=&\eta_{i_n}\mu_{i_n}.\\
{\bf{\hat t}}_{2i_n}\cdot{\mathbf\sigma}_{i_n}&=&\xi_{i_n}\mu_{i_n}.
\end{eqnarray}
These completes the definitions of Jordan-Wigner fermionisations used in this article.
\subsubsection{The gauge fixed Hamiltonian}
The Hamiltonian can be written in terms of the gauge invariant
fermions as,
\begin{equation}
\label{gfham}
\tilde H=J_x\sum_{\langle ij\rangle_x}i\eta_i\tilde u_{ij}\eta_j
+J_y\sum_{\langle ij\rangle_y}i\eta_i\tilde u_{ij}\eta_j
+J_z\sum_{\langle ij\rangle_z}i\eta_i\tilde u_{ij}\eta_j,
\end{equation}
where the gauge fixed Z$_2$ fields, $\tilde u_{ij}$ are,
\begin{eqnarray}
\nonumber
\tilde u_{ij}&=&i\xi_i\xi_j~~~~{\rm normal~bonds}, \nonumber\\
\nonumber
&=&1~~~~~~~~{\rm tangential~bonds~except}~(ij)=(i_1i_{N_S}), \nonumber\\
\tilde u_{i_1i_{N_S}}&=&{\cal S} \nonumber\\
&=&\prod_{n=1}^{N_S}~\left({\bf{\hat n}}_{i_m}\cdot{\mathbf\sigma}_{i_m}\right) \nonumber \\
&=&
~\left(u_{i_{1}i_{N_S}} u_{i_{N_S}i_{N_S-1}}
\dots u_{i_3 i_2}u_{i_2i_1}\right) \prod_{n=1}^{N_S} \eta_{i_n}.
\label{utildedef}
\end{eqnarray}
${\cal S}$ is a well known conserved quantity in the one-dimensional
applications where it corresponds to the total number of fermions modulo 2
and determines the boundary conditions on the fermions.
Thus the Jordan-Wigner transformation is equivalent to a gauge fixing
procedure where all the gauge fields on the tangential bonds (except one)
are set equal to 1. The choice of the Hamilton path amounts to a gauge
choice since it defines which of the bonds are tangential. It also
defines the sign in the definition of $u_{ij}$ for the normal bonds. In
Eq.~(\ref{utildedef}), the sign corresponds to a Hamilton path which
winds regularly in the ${\bf{\hat e}}_1$ direction as shown in
Fig (\ref{jwpt1}). Hereinafter all explicit computations will be with respect to
this path. A general algorithm to go to this gauge, which we will refer to
as the Jordan-Wigner gauge, is given in appendix \ref{gfalg}.
\begin{figure}[h!]
\center{\hbox{\epsfig{figure=hexagon.eps,height=1.7in,width=2.9in}}}
\caption{Jordan-Wigner path on a torus for $4\times4$ lattice. The numerics at the lattice sites describe how the Jordan-Wigner path traverses the lattice.}
\label{jwpt1}
\end{figure}
\subsubsection{The fermionic conserved quantities}
\label{frconsq}
All the gauge fixed Z$_2$ fields are conserved quantities. It is convenient
to express them in terms of the gauge fixed bond fermions on the normal bonds,
\begin{equation}
\chi_{ij}\equiv\frac{\xi_i+i\xi_j}{2},~~~
\chi^\dagger_{ij}\equiv\frac{\xi_i-i\xi_j}{2}.
\label{chidef}
\end{equation}
The conserved quantities are then the occupation numbers of the bond fermions,
\begin{equation}
\tilde u_{ij}=2~\chi^\dagger_{ij}\chi_{ij}-1.
\end{equation}
There are $N_p$ normal bonds and hence the bond fermion occupation numbers
form a set of $N_p$ conserved quantities. Thus along with ${\cal S}$,
we have $N_p+1$ conserved quantities consistent with the analysis in terms
of spin variables in section \ref{spconsq}. To see the meaning of ${\cal S}$,
it is convenient to define complex fermions in the matter sector from the
two $\xi$ fermions on every normal bond,
\begin{equation}
\label{psidef}
\psi_{ij}\equiv\frac{\eta_i+i\eta_j}{2},
~~~\psi^\dagger_{ij}\equiv\frac{\eta_i-i\eta_j}{2}.
\end{equation}
It can then be shown that,
\begin{eqnarray}
{\cal S}&=&(-1)^{\left( N_\psi+N_\chi+1\right)}, \nonumber\\
N_\psi&\equiv&\sum_{\rm z~bonds}\psi^\dagger_{ij}\psi_{ij}, \nonumber\\
N_\chi&\equiv&\sum_{\rm normal~bonds}\chi^\dagger_{ij}\chi_{ij}.
\end{eqnarray}
\subsubsection{From Kitaev gauge to Jordan-Wigner gauge}
\label{kjg}
We have explained that Jordan-Wigner gauge is a special realisation of Kitaev gauge where all the gauge fields residing on the tangential bonds are fixed to unity. One may wonder whether there exists a gauge transformation on the lattice which renders Eq.~(\ref{z2ham1}) to that of Eq.~(\ref{gfham}). Indeed there exists such a gauge transformations. Referring to the Jordan-Wigner path given in Fig. (\ref{jwpt1}) we do the following gauge transformations at a site `$n$'
\begin{eqnarray}
c_{n} \rightarrow \prod_{i=1,n-1} u_{i,i+1}c_{n}, ~ 1 <n <N.
\end{eqnarray}
In the above equation the various indices corresponds to the Jordan-Wigner path in the Fig. (\ref{jwpt1}). $u_{i,i+1}$s are the Z$_2$ gauge fields that would normally exists on the link joining site `$i$' and `$i+1$' if we apply the Majorana fermionisation as adopted by Kitaev. After implementing the above gauge tranformations
Z$_2$ gauge fields appear only on the normal bonds. The above gauge transformations yield a new
conserve quantity $\mathcal{S}^{\prime}$ which appear on the bond where the Jordan-Wigner end path meets.
The expression for $\mathcal{S}^{\prime}$ is,
\begin{eqnarray}
&&\mathcal{S}^{\prime}= \prod_{i=1,N-1} u_{i,i+1}.
\end{eqnarray}
The significance of $\mathcal{S}^{\prime}$ is very similar to $\mathcal{S}$ introduced in Eq.~(\ref{utildedef}).
\section{Four fold Degeneracy on a torus}
\label{gsd}
Since the gauge fields are static, the problem is reduced to one of
non-interacting fermions on a lattice. It is known \cite{kit1, lieb}
that the lowest fermionic ground state energy is obtained for the
flux free configuration, i.e. $B_p=1,~\forall p$. On the torus, there
are four gauge in equivalent configurations for every configurations of $B_p$ as argued in \ref{spconsq}. These
correspond to the four values of the gauge invariant conserved quantities
$W_1~{\rm and}~W_2$. First We examine these four different gauge field configurations for flux free sector and compute the corresponding fermionic ground state. Following this we show that this leads to four fold degeneracy of ground states in thermodynamic limit. Next we demonstrate explicitly how to obtain the four fold degeneracy for every configurations of fluxes. To this end we derive the required operators which enables us to obtain any one of the inequivalent gauge field configurations from the other for arbitrary flux configurations.
\subsection{Degenerate ground states on a Torus}
\label{gsd1}
~~~~To start with we briefly recapitulate the notions of Jordan-Wigner transformation and refer to Fig. (\ref{jwpt1}). The normal bonds are the ones that form the basis of the triangular lattice except for the $(0,n,\alpha)$ line. On this line the normal bonds are the ones
between $(0,n,a)~{\rm and}~(0,n+1,b))$. We choose the first site of the
path to be $i_1=(0,0,b)$. The four flux free configurations are then explained
as follows. To this end we write the exact Hamiltonian and $B_p$ for this particular realisation of Jordan-Wigner transformation. We divide the Hamiltonian in three parts, $H_{int}$ , $H_{bound}$ and $H_{end}$. $H_{int}$ includes all the internal bonds and $H_{bound}$ includes all the boundary bonds except one where the Jordan-Wigner end points meets. $H_{end}$ includes the interaction for the bond where the Jordan-Wigner end points meet each other. Similarly all the $B_{p}$'s are categorized in the above three different way. Below we write the various parts of the Hamiltonian and $B_p$'s.
\begin{eqnarray}
H_{int} &=& \sum_{m,n} iJ_{x} \eta^{a}_{m,n} \eta^{b}_{m+1,n+1} + \sum_{m,n} iJ_{y} \eta^{a}_{m,n} \eta^{b}_{m,n+1} \nonumber \\
&&
+ \sum_{m,n} iJ_{z} \tilde{u}_{m,n}\eta^{a}_{m,n} \eta^{b}_{m,n} \label{hint},
\end{eqnarray}
where $\tilde{u}_{m,n}$ is defined on each internal z-bonds.
\begin{eqnarray}
H_{bound} &=& \sum_{m,n} iJ_{y} \tilde{u}_{\substack{m,n\\m,n+1}} i\eta^{a}_{m,n}\eta^{b}_{m,n+1} \nonumber \\
&& +\sum_{m,n} J_{z} i \eta^{a}_{m,n} \eta^{b}_{m,n}. \label{hbound}
\end{eqnarray}
Where $\tilde{u}_{\substack {m,n\\ m,n+1}}$ is defined on each boundary y-bond.
The Hamiltonian for the end bond is given by,
\begin{eqnarray}
H_{end}&=& -{\cal S} i\eta^{a}_{M,N} \eta^{b}_{M,N}. \label{hend}
\end{eqnarray}
Now with the definition of $\psi$ fermion and $\chi$ fermion we get the equations (\ref{hint}), (\ref{hbound}) to be rewritten as,
\begin{eqnarray}
H_{int} &=& \sum_{m,n} J_{x}(\psi^{\dagger}_{m,n} +\psi_{m,n})(\psi^{\dagger}_{m+1,n+1} -\psi_{m+1,n+1}) \nonumber \\
& & +J_{y}(\psi^{\dagger}_{m,n} +\psi_{m,n})(\psi^{\dagger}_{m,n+1} -\psi_{m,n+1})\nonumber\\
& &+ J_{z}\tilde{u}_{m,n}(2\psi^{\dagger}_{m,n}\psi_{m,n}-1).
\end{eqnarray}
The Hamiltonian for the slanting bonds,
\begin{eqnarray}
H_{bound}&=& \sum_{m,n} J_{y}\tilde{u}_{\substack{m,n\\m,n+1}}(\psi^{\dagger}_{m,n} +\psi_{m,n})(\psi^{\dagger}_{m,n+1} -\psi_{m,n+1}) \nonumber \\
& &+ J_{z}(2\psi^{\dagger}_{m,n}\psi_{m,n}-1).
\end{eqnarray}
Lastly the Hamiltonian term for the end bond where the end points of the Jordan-Wigner path meet each other is given by,
\begin{equation}
H_{end}=-{ \cal S} J_{z} (2\psi^{\dagger}_{M,N}\psi_{M,N}-1).
\end{equation}
Following equations (\ref{utildedef}), (\ref{xidef}) and (\ref{psidef}), we rewrite the complete expressions for various conserved quantities appeared in the final form of the Hamiltonian. The conserve quantity $\tilde{u}_{m,n}$ defined on each internal z-link is given by,
\begin{equation}
\tilde{u}_{m,n}=(2\chi^{\dagger}_{m,n} \chi_{m,n}-1).
\end{equation}
Similarly the conserved quantity defined on each boundary y-bonds (which are labeled by the z-bonds it is connected with (i,e $\tilde{u}_{\substack{m,n\\m,n+1}}$))
is given by,
\begin{equation}
\tilde{u}_{\substack{m,n\\m,n+1}}= (2\chi^{\dagger}_{\substack{m,n\\m,n+1}}\chi_{\substack{m,n\\m,n+1}}-1 ).
\end{equation}
And the ${\cal S}$, for the Jordan-Wigner gauge, is given by,
\begin{equation}
{\cal S}= -(-1)^{MN+N_{\psi}+N_{\chi}}.
\end{equation}
If $ \mathcal{P}^{\mathcal G}$ and $ \mathcal{P}^{\mathcal M}$ denote the parity operators for the gauge fermions and the matter fermions respectively, then we can write $\mathcal{S}= -(-1)^{MN} \mathcal{P}^{\mathcal M}\mathcal{P}^{\mathcal G}$. From the Fig. \ref{jwpt1}, we see that a single hexagon always contains two normal bonds where each normal bond is associated with a conserved(static) Z$_2$ gauge field . $B_p$ for any plaquette is the product of these two conserved Z$_2$ gauge fields. Thus, $B_p = \tilde{u}_{ij} \tilde{u}_{kl}$ , where `$ij$' and `$kl$' are the normal bonds for the plaquette `$p$'. For the end plaquette where Jordan-Wigner path terminates $B_p$ is given by $B_p = - {\cal S}\tilde{u}_{ij} \tilde{u}_{kl}$. \\
Now we are in a position to show the ground state degeneracy in thermodynamic limit. To this end we explicitly write the four in-equivalent gauge field configurations corresponding the flux free configurations and write down the corresponding fermionic Hamiltonian. Finally we find the spectra for each of these four fermionic Hamiltonian.
\subsubsection {Choice 1}
Here the flux free configuration is obtained by making all the $\tilde{u}$'s to be 1 and ${\cal S}=-1$. The loop conserve quantities are having the following eigenvalue ,$W_{1}=1$ and $W_{2}=1 $. These particular choice makes the resulting fermionic Hamiltonian transitionally invariant and usual Periodic boundary condition in both the direction can be used to diagonalize the Hamiltonian. We will explicitly write the Hamiltonian in terms of complex fermion $\psi$. However one can equivalently work in terms of $\eta$, Majorana fermion representation. To keep this in mind we will continue to mention appropriate gauge transformations for $\eta$ fermions as well as $\psi$ fermions.
The complete translational invariant Hamiltonian is given by,
\begin{eqnarray}
H &=& \sum_{m,n} J_{x}(\psi^{\dagger}_{m,n} +\psi_{m,n})(\psi^{\dagger}_{m+1,n+1} -\psi_{m+1,n+1}) \nonumber \\
& & +J_{y}((\psi^{\dagger}_{m,n} +\psi_{m,n}))(\psi^{\dagger}_{m,n+1} -\psi_{m,n+1})\nonumber\\
& &+J_{z}(2\psi^{\dagger}_{m,n}\psi_{m,n}-1).
\end{eqnarray}
This is a manifestly p-wave superconducting Hamiltonian and can be easily diagonalized by going to the momentum space.
The constraint on the number of $\psi$ fermion becomes,
\begin{equation}
\prod (2\psi^{\dagger}_{m,n}\psi_{m,n}-1)=1,
\end{equation}
which implies that we are to fill up only the even number of $\psi$ fermions. We define the Fourier transform of the $\psi_{m,n}$ as given below,
\begin{equation}
\psi_{m,n}=\frac{1}{\sqrt{MN}}\sum_{p,q} e^{i\left(k_{1}m+k_{2}n\right)} \label{ftc1},
\end{equation}
where $k_{1}= 2\pi\frac{p}{M} $ and $k_{2}= 2\pi\frac{q}{N}$.
This is obtained by noticing the fact that we can write ${\vec{k}}=\frac{p}{M}{\vec{G_1}}+\frac{q}{N}{\vec G_2}$ where ${\vec G_{1/2}} $ are the reciprocal lattice vectors which are given by,
\begin{equation}
{\vec{G_{1}}}= \frac{4\pi}{\sqrt{3}}\left(\frac{\sqrt{3}}{2} \textbf{e}_{x} +\frac{1}{2} \textbf{e}_{y}\right) \,\,;\,\, {\vec{G_{2}}} = \frac{4\pi}{\sqrt{3}} \textbf{e}_{y}.
\end{equation}
Substituting this we get the resulting Hamiltonian in momentum space as,
\begin{eqnarray}
H &=&\sum_{k} (\epsilon_{k} \psi^{\dagger}_{k} \psi_{k} - \epsilon_{k} \psi_{-k} \psi^{\dagger}_{-k} +i \delta_{k} \psi^{\dagger}_{k} \psi^{\dagger}_{-k} -i \delta_{k} \psi_{-k} \psi_{k})\nonumber\\
&&+ \epsilon_{0,0} \psi^{\dagger}_{0,0}\psi_{0,0} + \epsilon_{\pi,0} \psi^{\dagger}_{\pi,0}\psi_{\pi,0}+\epsilon_{0,\pi} \psi^{\dagger}_{0,\pi}\psi_{0,\pi}\nonumber\\
& &+\sum_{k} \epsilon_{k} -MN J_z \label{1ckm}.
\end{eqnarray}
Where $\epsilon_{k}=2(J_x \cos k_x +J_y \cos k_y +J_z) $ and
$ \delta_k=2(J_x \sin k_x +J_y \sin k_y )$.
$k_{x}=\textbf{k}.\textbf{n}_x$, $k_{y}=\textbf{k}.\textbf{n}_y$ and
$\textbf{n}_{x,y}=\frac{1}{2}\textbf{e}_{x} \pm
\frac{\sqrt{3}}{2}{\textbf{e}_{y}}$ are unit vectors along x and y type bonds.
In Eq.~(\ref{1ckm}), the sum over `$k$' runs over first half of the Brillouin zone and does not
include the `$k$'-points $(\pi,0), (0,\pi),(0,0)$. The first line of the Hamiltonian
is diagonalize by the following transformations,
\begin{equation}
\left( \begin{array}{r}
\alpha_{k}\\
\beta_{k} \end{array}\right)=\left( \begin{array}{rr}
\rm cos \theta_{k} & -\rm i \;\rm sin\theta_{k} \\
\rm -i \; sin \theta_{k} & \rm cos \theta_{k} \\
\end{array} \right) \left( \begin{array}{r}
\psi_{k}\\
\psi^{\dagger}_{-k} \end{array}\right).
\end{equation}
Where $\cos2\theta_{k}=\epsilon_{k}/E_{k}$, with $E_{k}=\sqrt{\epsilon^2_k + \delta^2_k}$.
Then re-witting the Hamiltonian we get,
\begin{eqnarray}
H&=&\sum E_{k}(\alpha^{\dagger}_{k}\alpha_{k}-\beta^{\dagger}_{k}\beta_{k})\nonumber\\
&& +\epsilon_{0,0} \psi^{\dagger}_{0,0}\psi_{0,0} + \epsilon_{\pi,0} \psi^{\dagger}_{\pi,0}\psi_{\pi,0}+\epsilon_{0,\pi} \psi^{\dagger}_{0,\pi}\psi_{0,\pi} \nonumber\\
&&-\frac{1}{2}(\epsilon_{0,0}+\epsilon_{0,\pi}+\epsilon_{\pi,0}) +(\sum_{k^{\prime}}\frac{1}{2}\epsilon_{k^{\prime}}-N_z J_z).
\end{eqnarray}
Here sum over $k^{\prime}$ runs over full Brillouin zone. The last term in the parenthesis is always zero for Torus.
\subsubsection{Choice 2}
Here the flux free configuration is obtained by making all $\tilde{u}$'s -1 and ${\cal S}=-1$. The corresponding values of loop conserve quantity is given by,$W_1=-1$ and $W_2=-1$. To implement the Fourier transformation we do the following steps.
We make the following gauge transformation. $\eta^{b}_{M,n} = -\eta^{b} _{M,n}$ for all `$n$'. In terms of $\psi$ fermion the necessary gauge transformation is $\psi_{M,n}=-\psi^{\dagger}_{M,n}$ for all `$n$'. Then the Hamiltonian requires $\eta_{M+1,n}=-\eta_{1,n}$ and $\eta_{m,N+1}=\eta_{m,1}$(or alternatively $\psi_{M+1,n}=-\psi_{1,n}$ and $\psi_{m,N+1}=\psi_{m,1}$). This is equivalent to anti-periodic boundary Condition in $\bf e_{1}$ direction and periodic boundary condition in $\bf e_{2}$ direction. The necessary Fourier transform is defined with,
\begin{equation}
\psi_{m,n}=\frac{1}{\sqrt{MN}}\sum_{p,q} e^{i\left(k_{1}m+k_{2}n\right)},
\end{equation}
with $k_{1}= \frac{2\pi}{M}(m+\frac{1}{2}); \,\,\,\, k_{2}= 2\pi\frac{n}{N}$.
Substituting this in the Hamiltonian and diagonalizing straightforwardly we get,
\begin{eqnarray}
H&=&\sum_k E^{2}_{k}(\alpha^{\dagger}_{k}\alpha_{k}-\beta^{\dagger}_{k}\beta_{k}) +(\sum_{k^{\prime}}\frac{1}{2}\epsilon_{k^{\prime}}-N_z J_z) \label{2ckm}.
\end{eqnarray}
Here also `$k$' runs over first half of the Brillouin zone and the `$k^{\prime}$' runs over the full Brillouin zone. Note the absence of $(0,\pi),(\pi,0) \rm \, (0,0)$ mode. They do not appear here for this anti-periodic boundary condition. This will be true for the choices 3 and 4 also. Various parameters appearing in Eq.~(\ref{2ckm}) are given below,
\begin{eqnarray}
&&E^{2}_{k}=\sqrt{(\epsilon_{k}^{2}+\delta^{2}_{k})}, \nonumber \\
&& \epsilon_{k}=2(J_{x} \cos k_{x} +J_{y}\cos k_{y} -J_{z}), \nonumber \\
&& \delta_{k}=2(J_{x}\sin k_{x}+J_{y}\sin k_{y}).
\end{eqnarray}
\subsubsection{Choice 3}
For this case the flux free configuration is obtained by making $\tilde{u}_{M,n}=-1 \rm $ where `$n$' runs from 1 to $N-1$. All other $\tilde{u}$'s are 1 and ${\cal S}=1$. The loop conserve quantity $W_1$ takes value 1 and and $W_2$ -1. Similar to the previous case we need to do following gauge transformations in order to apply Fourier transform. We make $\eta^{b}_{m,N} = -\eta^{b} _{m,N}$ for all `$m$'. In terms of $\psi$ fermion the necessary gauge transformation is $ \psi_{m,N}=-\psi^{\dagger}_{m,N}$ for all `$m$. The resulting Hamiltonian requires $\eta_{m,N+1}=-\eta_{m,1}$ and $\eta_{M+1,n}=\eta_{1,n}$( $\psi_{m,N+1}=-\psi_{m,1}$ and $\psi_{M+1,n}=\psi_{1,n}$). This indicates anti-periodic boundary condition in $\bf e_{2}$ direction and periodic boundary condition in $\bf e_{1}$ direction. The resulting Fourier transform is defined with,
\begin{equation}
k_{1}= 2\pi\frac{m}{M}; \,\,\,\,\, k_{2}= \frac{2\pi}{N}(n+\frac{1}{2}).
\end{equation}
The resulting Hamiltonian in `k' space is similar to Eq.~(\ref{2ckm}) with $\epsilon_{k}=2(J_{x} \cos k_{x} +J_{y}\cos k_{y} +J_{z})$.
\subsubsection{Choice 4}
For this choices we need, $\tilde{u}_{M,n}=1 \rm$ where `$n$' runs from 1 to $N-1$ . All other $\tilde{u}$'s are -1 and ${\cal S}=1$. $W_1=-1$ and $W_2=1$. In this case one requires combined gauge transformation mentioned for the choices 2 and 3. This makes the Hamiltonian anti periodic in both the directions. The required Fourier transform is defined with,
\begin{equation}
k_{1}= \frac{2\pi}{M}(m+\frac{1}{2}) ;\,\,\,\, k_{2}= \frac{2\pi}{N}(n+\frac{1}{2}).
\end{equation}
Proceeding as before we get exactly Eq.~(\ref{2ckm}) with identical expression for $\epsilon_{k}$.
\subsubsection{Ground State Energy in thermodynamic limit}
To get the ground state energy for the gauge choices 2,3 and 4, one fills up the negative energy states consistent with the boundary condition (i.e, to satisfy the
constraint ${\cal S}$ which restricts the total number of particles (odd or even number) to be taken). In the limit of $M ,N \rightarrow \infty $ all the above four choices, the ground state energy is obtained as,
\begin{equation}
E_{G}= \frac{\sqrt{3}}{16\pi^{2}}\int_{BZ} E(k_{x},k_{y})dk_{1}dk_{2},
\end{equation}
with $E(k_{x},k_{y})$ is defined before. The appearance of a `-' sign in the expression of $\epsilon_{k}$ for choices 2 and 4 can be accounted for by shifting the $k_1$ integral to $\pi - k_1$. Thus it is clear that in thermodynamic limit ground state has four fold degeneracy.
\subsection{Four fold Topological degeneracy for any eigenstate}
\label{tpg}
In the preceding section, we have shown that four degenerate ground states are
characterized by four different topologically distinct gauge field
configurations. From the section \ref{z2gtkm}, we infer that every flux
configuration is characterized by such four topologically distinct gauge field
configurations. This leads to a four-fold degeneracy for every eigenstate,
including the ground state, in the thermodynamic limit.
We now demonstrate this explicitly. Our proof is similar to that of
Wen and Niu \cite{wen-niu} which shows the topological degeneracy of
fractional quantum Hall states on a torus. We construct two operators which
we call $V_1$ and $V_2$, that act on states with a given gauge field
configuration and produce states with a different gauge field configuration
without changing the values of the flux operators, $B_p$. They however change
the values of the Wilson loop operators, $W_{1,2}$. These two operators are
therefore the generators of the so called ``large gauge transformations".
There are four topologically different sectors of gauge field configurations
corresponding to $W_i=\pm 1$.
We further show that only effect of the large gauge transformations on the
matter sector is to change the boundary conditions of the Majorana fermions
from periodic to anti-periodic or vice versa. Thus the energy eigenvalues
only change by $\sim 1/L$, where $L$ is the length of the torus. The
eigenstates in the four sectors (related by the action of $V_i$) are therefore
degenerate in the thermodynamic limit.
The four operators $V_i,~W_i$ characterise the topological degeneracy.
We show that $V_i$ and $W_i$ satisfy the following algebra,
\begin{eqnarray}
\{V_1,V_2\}=0,~~ \{V_i,W_i\}=0,~~ [V_i,W_j]_{i\ne j}=0.\label{cmtr}
\end{eqnarray}
We can then construct four operators, $T_a,~a=1,\dots,4$,
\begin{eqnarray}
\label{tadef}
T_1=V_1 W_1, ~~ T_2=V_2 W_2,~~ T_3=V_1, ~~ T_4=V_2.
\end{eqnarray}
These satisfy the Clifford algebra,
\begin{equation}
\label{cliff}
\{T_a,T_b\}=2\delta_{ab}
\end{equation}
Thus we show that the four-fold topological degeneracy on the torus
is characterised by the 4-dimensional Clifford algebra.
We will first write down and discuss the expressions for $V_1$ and $V_2$
for the 32 site system illustrated in Fig. \ref{jwpt1}. The construction
easily generalises for any even-even lattice.
\begin{eqnarray}
\label{visdef}
V_1&=&\sigma^z_1 \sigma^z_3 \sigma^z_5 \sigma^z_7\\
V_2&=&\sigma^y_1\sigma^y_{16}\sigma^y_{17}\sigma^y_{32}
\prod^7_{i=2}\sigma^z_i\prod^{23}_{i=18}\sigma^z_i
\end{eqnarray}
The two Wilson loops for this lattice are,
\begin{eqnarray}
\label{w1s}
W_1&=&
\sigma^y_1\sigma^y_8\sigma^y_9\sigma^y_{16}\sigma^y_{17}
\sigma^y_{24}\sigma^y_{25}\sigma^y_{32}\\
W_2&=&\prod_{i=1}^8\sigma^z_i
\end{eqnarray}
It can be verified that the above constructions satisfy the algebra
in equation (\ref{cmtr}) and hence the topological operators defined
in equation (\ref{tadef}) satisfy the Clifford algebra (\ref{cliff}).
It can also be verified that $V_i$ commute with all the $B_p$'s.
Now consider simultaneous eigenstates of the Wilson loop operators,
\begin{equation}
\label{westates}
W_i\vert w_1,w_2\rangle=w_i\vert w_1,w_2\rangle~,~~~~~~~~w_i=\pm 1
\end{equation}
The algebra in equation (\ref{cmtr}) implies,
\begin{equation}
\label{vonwes}
V_1\vert w_1,w_2\rangle=\vert -w_1,w_2\rangle,~~
V_2\vert w_1,w_2\rangle=\vert w_1,-w_2\rangle
\end{equation}
Thus we have shown that $V_i$ are the generators of large gauge
transformations.
We next consider their action on the hamiltonian.
\begin{equation}
\label{vonh}
V_1HV_1^{-1}=H^1,~~~~~~V_2HV_2^{-1}=H^2
\end{equation}
where $H^1$ is the same as $H$, except that the bonds on
one non-trivial loop in the ${\bf e}_1$ direction have changed sign,
namely the loop $(1,2,3,4,5,6,7,8,1)$. In $H^2$, a line of parallel
bonds in the ${\bf e}_1+{\bf e}_2$ direction have changed sign. Namely,
the bonds $(7-8),~(31-32),~(23-24)$ and $(16-17)$. We will now write down
the operators for a general even-even lattice and then show that the
transformed hamiltonians are degenerate in the thermodynamic limit. We
will show that in the fermionised theory, these changes of sign can be
absorbed into the single particle eigenfunctions of the Majorana fermions
and change the single particle energy eigenvalues by $\sim 1/L$. Thus the
energies do not change in the thermodynamic limit, making every
many body eigenstate four-fold degenerate.
The general expressions for $V_{1(2)}$ are given by
\begin{eqnarray}
&&V_1= \prod^{M}_{m=1} \sigma^{b,z}_{m,0},
\nonumber \\
&&V_2= \prod^{N/2} _{n=1} \sigma^{a,y}_{0,2n} \sigma^{b,y}_{0,2n}
\prod^{m=1,M-1}_{n=1,N/2} \sigma^{a,z}_{m,2n-1} \sigma^{b,z}_{m,2n-2} \label{v1v2}.
\end{eqnarray}
It can be verified that these constructions satisfy the algebra in equation
(\ref{cmtr}) and also commute with all the $B_p$'s. In $H^1$, the bonds
$(m,0,b)-(m+1,1,a) ~ {\rm{and}}~ (m,1,a)-(m,0,b)$ change sign and in $H^2$, the bonds $(M-1,n,b)-(0,n+1,a)$
change sign.
In the fermionised theory, the single particle eigenfunctions satisfy the
equation,
\begin{equation}
\label{spe1}
\sum_jiA_{ij}\phi^n_j=\epsilon^n\phi^n_i
\end{equation}
where $A_{ij}$ is an antisymmetric matrix coupling the nearest neighbours
of the honeycomb lattice. The eigenvalues come in pairs and we denote
$(\phi^n)^*=\phi^{-n}$, $\epsilon^{-n}=-\epsilon^n$. $n$ will then go
from $1,\dots,NM$. The Hamiltonian is diagonal in terms of the complex
fermions defined by,
\begin{eqnarray}
\label{alphadef}
\alpha_n&=&\sum_i\phi^n_{ia}\eta_i^a+\phi^n_{ib}\eta_i^b\\
\label{diagham}
H&=&\sum_n\epsilon^n\left(2\alpha^\dagger_n\alpha_n-1\right)
\end{eqnarray}
We now make the transformation,
\begin{equation}
\label{philgt1}
\phi^{n\prime}_{nm,a}= e^{in\pi/N}\phi^n_{nm,a}
\end{equation}
Equation (\ref{spe1}) then gets written as,
\begin{equation}
\label{spe2}
\sum_j A_{ij} \phi^n_j=\sum_j\left(iA^1_{ij}+i\delta A_{ij}\right)\phi^{n\prime}_i
=\epsilon^n\phi^{n\prime}_i
\end{equation}
where $A^1$ is the antisymmetric matrix corresponding to $H^1$ and
$\delta A_{ij}\propto 1/N$ when $N$ is very large. Thus the single
particle energy eigenvalues of $A$ and $A^1$ are identical in the
thermodynamic limit when $N \rightarrow \infty$. The spectrum of $H$ and $H^1$ are also therefore
identical. The mapping of the eigenvalues of $H$ and $H^2$ can be similarly
shown using the transformation,
\begin{equation}
\label{philgt2}
\phi^{n\prime}_{nm,a}= e^{im\pi/M}\phi^n_{nm,a}
\end{equation}
We can also write equation (\ref{v1v2}) in terms of the gauge invariant
fermions. $V_1$ or $V_2$ can be written as $V_i=V^{\mathcal M}_i V^{\mathcal G}_i$ ($i=1,2$). As the
notation suggests, $V^{\mathcal M}_i$'s and $V^{\mathcal G}_i$'s are composed of operators which
belong to the matter sector and the gauge sector respectively. They are
given by,
\begin{eqnarray}
&&V^{\mathcal G}_1=\prod \xi^a_{m,0}, ~ m=1, M-1, \nonumber \\
&&V^{\mathcal M}_1=\prod \eta^{a}_{m,0}, ~m=1,M, \nonumber \\
&&V^{\mathcal G}_2= \prod^{N/2} _{n=1} \xi^{a}_{0,2n} \xi^{b}_{0,2n}
\prod^{j=1,N/2}_{k=2,N} \xi^{a}_{m,2n-1} \xi^{b}_{m,2n-2}, \nonumber \\
&&V^{\mathcal M}_2= \prod^{N/2} _{n=1} \eta^{a}_{0,2n} \eta^{b}_{0,2n}
\prod^{m=1,M-1}_{n=2,N/2} \eta^{a}_{m,2n-1} \eta^{b}_{m,2n-2}.
\end{eqnarray}
From the above expressions we can easily find the commutation relations of parity operator for gauge fermions and the matter fermions. We find that,
\begin{eqnarray}
&&\{\mathcal{P}^{G}, V_1\}=0,~[\mathcal{P}^{\mathcal G}, V_2]=0,~[\mathcal{P}^{\mathcal M}, V_1]=0,~[\mathcal{P}^{\mathcal M}, V_2]=0. \nonumber \\
&&
\end{eqnarray}
The fact that the parity of the matter fermions are conserved is consistent with the fourfold degeneracy discussed here and in (\ref{gsd1}).
\section{Discussion}
\label{dis}
In this paper we have discussed many important aspects of Kitaev model. We
have shown how the $SU2$ gauge contained in the Kitaev model and explained
explicitly how this $SU2$ gauge symmetry is reduced to Z$_2$ gauge symmetry.
We have solved the Kitaev model using Jordan-Wigner method in a general way.
Though Jordan-Wigner transformations has been used earlier to solve for Kitaev
model, our formalism reveals many new features, for example, our definition of
J-W transformation is applied to a torus. We focused on the gauge field
contents of the Kitaev model and the issue of topological degeneracy. We showed
that ground state is four fold degenerate on the torus {\em in both phases}.
While it indicates non-trivial topological order, the ground state degeneracy
does not distinguish between the gapless and gapped phases. Finally we have
shown the equivalence between the fermionised Hamiltonian obtained in Kitaev
gauge and J-W gauge. Lastly we have constructed four mutually anti-commuting
operators on a torus to illustrate explicitely the four fold degeneracy for
every eigenstate. Our analysis reveals that Jordan-Wigner analysis can be
used even in quantum spin liquid problems, as a general method, to bring out
non trivial gauge field content, thereby providing an alternative method in
resonating valence bond theories.
\section{Acknowledgements}
As we were completing our manuscript, we learned of a very recent paper by
F. J. Burnell and Chetan Nayak (Ref. \cite{nayak}), which has some overlap with the present paper. GB thanks F. J. Burnell for
reference[40]. This research was partly supported by Perimeter Institute for Theoretical Physics.
|
\section{Introduction}
The \textit{water-wave problem} concerns the gravity-driven flow of a
perfect fluid of unit density; the flow is bounded below by a rigid
horizontal bottom $\{Y=-d\}$ and above by a free surface
$\{Y=\eta(X,t)\}$, where $\eta$ depends upon the horizontal spatial
coordinate $X$ and time $t$. \textit{Steady waves} are waves which
propagate from left to right with constant speed $c$ and without
change of shape, so that $\eta(X,t)=\eta(X-ct)$. \textit{Solitary
waves} are steady waves which have the property that
$\eta(X-ct)\rightarrow 0$ as $X-ct\rightarrow \pm\infty$. We consider
in this paper the particle trajectories in a fluid as a solitary wave propagates on the free surface, assuming
that the flow admits a negative
vorticity decreasing with depth.
There have been a series of works concentrating on the study of
solitary waves, in the setting of both irrotational flows
\cite{AT,BBB,CraigS, CE,CEH} etc.,
and rotational flows which become active only in the last few years.
One of the interests in the above works is the description of
individual
particle path in the fluid. In \textit{irrotational} flow, particle
paths underneath a solitary wave are investigated
in \cite{Cqam,CE}
, both for the smooth solitary wave and the solitary wave of greatest
height for which the crest is a stagnation point. It was shown in \cite{CE} that in an irrotational solitary water wave, each particle is
transported in the wave direction but slower than the wave speed; as the solitary wave propagates,
all particles located ahead of the wave crest are lifted while
those behind have a downward motion, with the particle trajectory having asymptotically the same height above the flat bed. For \textit{rotational }flow, some results on particle paths under \textit{periodic} waves are obtained (cf. \cite{Ejde, Enonlinearity, Delia,WaJDE} for instance), following a series of works on the corresponding results in irrotational case (see \cite{Cinvention,CW2,HenryImrn} etc).
Recently, rigorous existence results on small-amplitude solitary
waves with arbitrary vorticity distribution were obtained in \cite{HurSolitary} and in
\cite{GW}, using generalized implicit function theorem of
Nash-Moser type and spatial dynamics method, respectively. The solitary waves established in \cite{HurSolitary,GW} are of elevation and decays exponentially to a horizontal laminar flow far up- and downstream. The study
of solitary waves of large-amplitude with an arbitrary distribution of vorticity remains open. So this arises the question of particle paths in a rotational small-amplitude solitary wave.
Following the pattern in \cite{CE} for irrotational case, we
prove in this work the
corresponding results on
particle paths in rotational solitary waves by using the properties of solutions obtained in \cite{HurSolitary}. We consider in this work only solitary waves with negative vorticity, however with modifications our arguments can be applied also for positive vorticity.
Precisely we show that if the
vorticity is negative and increasing from bottom to surface of the
flow, and if the underlying current is favorable, i.e., is moving throughout the fluid in
the same direction as the wave,
then as time goes on the particle moves similarly as in the
irrotational case \cite{CE}. We also show that if the underlying current is not
favorable, then in solitary waves with
sufficiently small amplitude, some particles above the flat bed move
to the opposite direction of wave propagation along a path with a single loop or a single hump. Note that this single-loop kind of
path does not exist in the irrotational case. In \cite{HurSolitary}, as in most of
the works on waves with vorticity, the author
considered only waves that are not near breaking or
stagnation, i.e., the speed of an individual fluid particle is far
less than that of the wave itself throughout the fluid domain. We
consider only such waves as well. We do not consider the case of wave
with stagnation which is however studied in \cite{CE} in the
irrotational case.
The paper is organized as follows. In Section \ref{sec p} we present
the mathematical formulation for the solitary waves and recall from
\cite{HurSolitary} some useful properties of the established solitary wave solutions.
Section \ref{sec v} contains some conclusions on the vertical and horizontal velocity that are relevant for
our purposes. The main result is presented and proved in Section
\ref{sec r}. In the final section we give two examples which lie in our settings.
\section{Preliminaries}\label{sec p}
We first describe the governing equations for rotational solitary water waves and then recall some properties available on their solitary wave solutions.
\subsection{The governing equations for rotational solitary water waves.}
Choose Cartesian coordinates $(X,Y)$ so that the horizontal $X$-axis is in the direction of wave propagation and the $Y$-axis points upwards. Consider steady waves traveling at constant speed $c>0$. In the frame of reference moving with the wave, which is equivalent to the change of variables $(X-ct,Y)\mapsto (x,y)$, we use
\[
\Omega_{\eta}=\{(x,y)\in \mathbb{R}^2: -d<y<\eta(x)\},\quad 0 < d <\infty,
\]
to denote the stationary fluid domain and
$(u(x,y),v(x,y))$ to denote denote the velocity field, and define the \textit{stream function} $\psi(x,y)$ by $\psi(x,\eta(x))=0$ and
\begin{eqnarray}\label{def psi}
\psi_y=u-c,\quad \psi_x=-v.
\end{eqnarray}
Consider also only waves that are not near breaking or stagnation, so that $\psi_y(x,y)\leq -\delta<0$ in $\bar{\Omega}_{\eta}$ for some $\delta>0$, which implies that the vorticity $\omega=v_x-u_y$ is globally a function of the stream function $\psi$, denoted by $\gamma(\psi)$; see \cite{CW}.
The solitary-wave problem is then, for given $p_0<0$ and $\gamma\in C^1([0,-p_0]; \mathbb{R})$, to find a real parameter $\lambda$, a domain $\Omega_{\eta}$ and a function $\psi\in C^2(\bar{\Omega}_\eta)$ such that
\begin{eqnarray
\psi_y<0,\quad (x,y)\in \bar{\Omega}_{\eta},\label{psi y negative}\\
\triangle\psi=-\gamma(\psi),\quad (x,y)\in \Omega_{\eta},\label{equ psi}\\
\abs{\nabla \psi}^2+2gy=\lambda,\quad y=\eta(x),\label{Bernouli}\\
\psi=0,\quad y=\eta(x),\label{psi eta}\\
\psi=-p_0, \quad y=-d,\label{psi -d}\\
\eta(x)\rightarrow 0 \quad {\rm as} \abs{x}\rightarrow \infty,\label{eta to 0}\\
\psi_x(x,y)\rightarrow 0 \quad {\rm as~~} \abs{x}\rightarrow \infty ~~{\rm uniformly ~~ for~~} y.\label{psi x to 0}
\end{eqnarray}
Here $g>0$ is the gravitational constant of acceleration,
\begin{eqnarray*}
p_0=\int_{-d}^{\eta(x)}\psi_y(x,y)~dy<0
\end{eqnarray*}
is the relative mass flux (independent of $x$), and the boundary conditions \reff{eta to 0} and \reff{psi x to 0} express that the wave profile approaches a constant level of depth
and the flow is almost horizontal in the far field, respectively.
Moreover we require that the nontrivial solitary wave is of positive elevation, i.e., $\eta(x)>0$ for all $x\in\mathbb{R}$. It is therefore symmetric about its single crest and admits a strictly monotone wave profile on either side of this crest (see \cite {HurSymmetry}). Assuming the wave crest is located at $x=0$, the solitary-wave problem \reff{psi y negative}-\reff{psi x to 0} is thus supplemented with the symmetry and monotonicity conditions
\begin{eqnarray}\label{symmetry}
\psi(-x,y)=\psi(x,y),\quad \eta(-x)=\eta(x),\quad {\rm and}~~ \eta'(x)<0~~{\rm for ~~}x>0.
\end{eqnarray}
We refer to \cite{HurSolitary, CW} for more details on the derivation of the system \reff{psi y negative}-\reff{psi x to 0}.
\subsection{Rotational solitary water waves.} We collect some properties of the solitary wave solutions established in \cite{HurSolitary}. Let
\begin{eqnarray}\label{def Gamma}
\Gamma(p)=\int_{0}^{p}\gamma(-s)ds \quad {\rm and} ~~\Gamma_{\min}=\min_{p\in[p_0,0]}\Gamma(p)\leq 0.
\end{eqnarray}
Given $p_0<0$ and $\gamma\in C^1([0,-p_0]; \mathbb{R})$, for each $\lambda\in(-2\Gamma_{\min},\infty)$ the system \reff{psi y negative}-\reff{psi x to 0} admits a trivial solution pair $(\eta(x),\Psi(y))$ defined on $\bar\Omega_0$, where $\eta(x)\equiv0$, the stream function $\Psi(y)$ is $x$-independent and is the inverse of the function
\begin{eqnarray*}
y(\Psi)=\int_{p_0}^{-\Psi}\frac{d p}{\sqrt{\lambda+2\Gamma(p)}}-d,
\end{eqnarray*}
and the fluid domain
\begin{eqnarray*}
\Omega_0=\{(x,y)\in\mathbb{R}^2: -d<y<0\},\quad d=\int_{p_0}^{0}\frac{d p}{\sqrt{\lambda+2\Gamma(p)}}.
\end{eqnarray*}
The corresponding relative horizontal velocity and vertical velocity are thus given by
\begin{eqnarray}\label{U V}
U(y)-c=\Psi_y(y)=-\sqrt{\lambda+2\Gamma(-\Psi(y))},\quad V(x,y)=-\Psi_x(y)\equiv 0.
\end{eqnarray}
Note that $\Psi(0)=0$ and $\Psi(-d)=-p_0$. Thus
\begin{eqnarray*}
U(0)=c-\sqrt{\lambda}\quad {\rm and}\quad U(-d)=c-\sqrt{\lambda+2\Gamma(p_0)}.
\end{eqnarray*}
Throughout this paper, we adopt the terminology in \cite{CW2} to say this underlying trivial flow is \textit{favorable} if $U(y)\geq 0$ for all $y\in[-d,0]$, is \textit{adverse} if $U(y)<0$ for all $y\in[-d,0]$, and is \textit{mixed} if $U(y)$ changes sign. Note that favorable flow moves in the same direction as the wave propagation (i.e., to the right), while adverse flow moves in the opposite direction of the wave propagation.
To ensure the existence of nontrivial small amplitude solitary waves, the parameter $\lambda$ must be chosen to satisfy $\lambda>\lambda_c$ but close to $\lambda_c$, where $\lambda_c>-2\Gamma_{\min}$ is the unique solution of
\begin{eqnarray}\label{def lambda c}
\int_{p_0}^0\frac{dp}{\inner{\lambda_c+2\Gamma(p)}^{3/2}}=\frac{1}{g}.
\end{eqnarray}
For each such a given $\lambda$ and for given $p_0<0$, it was shown in \cite{HurSolitary} that there exists a nontrivial small amplitude solitary-wave solution pair $(\eta(x),\psi(x,y))$ to \reff{psi y negative}-\reff{psi x to 0} defined on $\bar\Omega_\eta$, with $\eta(x)$ satisfying
\begin{eqnarray}\label{eta small}
\abs{\eta(x)}+\abs{\eta'(x)}+\abs{\eta''(x)}\leq (\lambda-\lambda_c)r \quad {\rm for ~~all~~} x\in\mathbb{R}
\end{eqnarray}
and for some constant $r>0$ independent of $\lambda$, and the corresponding horizontal velocity satisfying the following properties:
\begin{enumerate}[~~~~(P1)]
\item \label{u to 0} $u(x,\eta(x))\rightarrow c-\sqrt{\lambda}=U(0)$, \quad {\rm as~~}$\abs{x}\rightarrow \infty $;
\item \label{u to U} $u(x,y)\rightarrow U(y)$ \quad {\rm as~~}$\abs{x}\rightarrow \infty$ ~~{\rm uniformly~~for~~} $y$.
\end{enumerate}
The first property is obvious since by \reff{Bernouli}, \reff{eta to 0} and \reff{psi x to 0}, we have $(u(x,\eta(x))-c)^2=\psi^2_y(x,\eta(x))\rightarrow \lambda$ as $\abs{x}\rightarrow \infty$, which equivalently gives (P\ref{u to 0})
as we have assumed $u(x,y)-c=\psi_y(x,y)<0$ in $\bar\Omega_\eta$, while the
property (P\ref{u to U}) can be deduced from the construction of solitary wave solutions; see \cite{HurSolitary}. Indeed for $\lambda=\lambda_c+\varepsilon$ with $\varepsilon>0$, there exists a function $w^{\varepsilon}(q,p)$, whose derivatives with respect to $(q,p)$ up to order 2 tend to 0 uniformly for $p$ as $\abs{q}\rightarrow \infty$, such that the horizontal velocity is determined by
\begin{eqnarray*
u(x,y)=c-\frac{1}{\inner{\lambda+2\Gamma(-\psi(x,y)}^{-1/2}+\varepsilon w^{\varepsilon}_p(\sqrt{\varepsilon}x,-\psi(x,y))},
\end{eqnarray*}
where $w^{\varepsilon}_p(q,p)$ denotes differentiation in
the $p$-variable.
We conclude this section by recalling some properties of streamlines. Due to $\psi_y<0$ throughout $\bar\Omega_\eta$, we have that for all $p\in[-p_0,0]$ the streamline
\[
\{(x,y): \psi(x,y)=-p\}
\]
is a smooth curve $y=\sigma_p(x)$. Note that
\begin{eqnarray}\label{sigma}
\sigma_0(x)=\eta(x),\quad \sigma_{p_0}(x)=-d\quad {\rm and ~~}\sigma_p'(x)=-\frac{\psi_x(x,\sigma_p(x))}{\psi_y(x,\sigma_p(x))}.
\end{eqnarray}
Observing the fact that $\frac{1}{\abs{\psi_y}}=\frac{1}{c-u}$ is bounded and that $\psi_x(x,y)\rightarrow 0$ uniformly for $y$ as $\abs{x}\rightarrow \infty$, we deduce that for each fixed $y_0\in[0,\eta(0)]$ the streamline $y=\sigma_p(x)$ with $p=-\psi(0,y_0)$, passing through the point $(0,y_0)$, has an asymptote $y=l(y_0)$ as $\abs{x}\rightarrow\infty$, with $l(\eta(0))=0$ and $l(-d)=-d$.
\section{Vertical and horizontal velocity}\label{sec v}
We first divide the fluid domain $\Omega_\eta$ into two components
\[
\Omega_-=\{(x,y)\in\mathbb{R}^2: x<0, 0<y<\eta(x)\} \quad {\rm and~~~} \Omega_+=\{(x,y)\in\mathbb{R}^2: x>0, 0<y<\eta(x)\},
\]
and denote their boundaries by
\[
S_-=\{(x,y)\in\mathbb{R}^2: x<0, y=\eta(x)\}, \quad B_-=\{(x,y)\in\mathbb{R}^2: x<0, y=-d\},
\]
respectively
\[
S_+=\{(x,y)\in\mathbb{R}^2: x>0, y=\eta(x)\}, \quad B_+=\{(x,y)\in\mathbb{R}^2: x>0, y=-d\}.
\]
\begin{lemma}\label{lem vertical}
Suppose that $\gamma'(p)\leq 0$ for all $p\in[0,\abs{p_0}]$. Then
\begin{enumerate}[(a)]
\item \label{v=0} $v(x,-d)=0$ for all $x\in\mathbb{R}$, and $v(0,y)=0$ for $y\in [-d,\eta(0)]$ ;
\item\label{lem vertical 2} $v(x,y)<0$ if $(x,y)\in \Omega_-\cup S_-$, and $v(x,y)>0$ if $(x,y)\in \Omega_+\cup S_+$;
\item\label{lem vertical 3} $v_y(x,-d)<0$ if $x<0$, and $v_y(x,-d)>0$ if $x>0$;
\item \label{lem vertical 4} $v_x(0,y)>0$ for $y\in(-d,\eta(0))$.
\end{enumerate}
\end{lemma}
\begin{proof}
Since $v=-\psi_x$, the first result follows from \reff{psi -d} and the symmetry property $\psi(-x,y)=\psi(x,y)$ in \reff{symmetry}.
Next we only prove the lemma for $x>0$ and the results for $x<0$ can be proved similarly. Differentiating \reff{psi eta} with respect to $x$ gives $v=-\psi_x=\psi_y\eta'(x)$, from which follows $v>0$ for $(x,y)\in S_+$ as $\psi_y<0$ and $\eta'(x)<0$ for $x>0$ in view of \reff{psi y negative} and \reff{symmetry}. To prove $v>0$ in $\Omega_+$, we assume first on the contrary that there exists a point $(x_0,y_0)\in\Omega_+$ such that $v(x_0,y_0)=-\varepsilon<0$. Then we can find a bounded domain
\[
\Omega_{+,k}=\{(x,y)\in\mathbb{R}: 0<x<k,-d<y<\eta(x)\},\quad k\in\mathbb{R}^+,
\]
such that $(x_0,y_0)\in\Omega_{+,k}$. Moreover in view of \reff{psi x to 0} we can choose $k$ sufficiently large so that $v(k,y)> -\varepsilon$ for $y\in[-d,\eta(k)]$. This means that $v$ attains its minimum at the interior point $(x_0,y_0)$ of $ \Omega_{+,k}$, which contradicts to the strong maximum principle applied to $v$ on the domain $\bar\Omega_{+,k}$, as $v$ satisfies
$\triangle v+\gamma'(\psi)v=0$ with $\gamma'(p)\leq 0$ by differentiating \reff{equ psi}. Therefore we have $v\geq 0$ in $\Omega_+$. If $v=0$ at a point $(x_0,y_0)$ of $\Omega_+$. Again we can choose a bounded domain $\Omega_{+,\hat k}$ containing $(x_0,y_0)$. Then $v\geq0$ on the boundary of $\Omega_{+,\hat k}$. We can thus apply the strong maximum principle on $\Omega_{+,\hat k}$ once more to conclude that $v\equiv0$ on $\bar\Omega_{+,\hat k}$, which contradicts $v>0$ on the half surface $S_+$. This proves $v>0$ in $\Omega_+$.
Since $v$ attains its minimum in $\bar\Omega_+$ on the half bed $B_+$ and on the crest line $x=0$, Hopf's maximum principle ensures that $v_y>0$ on $B_+$ and $v_x>0$ on $\{(0,y); -d<y<\eta(0)\}$, completing the proof.
\end{proof}
The above lemma combined with \reff{sigma} and the fact that $\psi_y<0$ gives
\begin{corollary}\label{cor streamline}
The streamline $y=\sigma_p(x)$ with $p\in(p_0,0]$ satisfies that $\sigma_p'(x)>0$ for $x<0$ and $\sigma_p'(x)<0$ for $x>0$.
\end{corollary}
\begin{lemma}\label{lem horizontal
Suppose that $\gamma(p),\gamma'(p)\leq0$ for all $p\in [0,\abs{p_0}]$, that $0<\lambda-\lambda_c<\varepsilon$ with $\varepsilon$ small such that nontrivial solitary wave exists, and that $c\geq\sqrt{\lambda+2\Gamma(p_0)}$. Then $u(x,y)>0$ for $x\in \bar\Omega_\eta$.
\end{lemma}
\begin{proof}
We assume throughout this proof that $\gamma(p)\not\equiv 0$ for $p\in [0,\abs{p_0}]$, since we already know that for $\gamma(p)\equiv 0$, i.e., the irrotational case, $u(x,y)>0$ in $\bar \Omega_\eta$; see \cite{CE}. Recall that $U(y)$ is the horizontal velocity of the trivial laminar flow. Thus $U'(y)=\Psi_{yy}(x,y)=-\gamma(\Psi(y))\geq0$, and there exists an interval $I$ such that the strict inequality holds for $y\in I$ since $\gamma(p)\not\equiv 0$ by assumption. This combined with \reff{U V} and the assumption $c\geq\sqrt{\lambda+2\Gamma(p_0)}$ gives
\begin{eqnarray*}
U(y)\geq U(-d)=c-\sqrt{\lambda+2\Gamma(p_0)}\geq 0~~{\rm for} ~~y\in[-d,0],~~~{\rm and }~~~~U(0)>0.
\end{eqnarray*}
It was proved in \cite{EV} that if $\gamma(p)\leq0$ for all $p\in [0,\abs{p_0}]$, then
\begin{eqnarray}\label{u mono s}
\frac{d}{dx}u(x,\eta(x))\geq 0~~~~{\rm for~~} x<0,~~~~{\rm and} ~~~~\frac{d}{dx}u(x,\eta(x))\leq 0~~{\rm for~~} x>0.
\end{eqnarray}
In other words, along the free surface $u$ increases from $x=-\infty$ to the crest $x=0$, and thereafter it is decreasing. On the bottom $B=\{(x,y)\in\mathbb{R}^2: y=-d\}$, we have, in view of $u_x=-v_y=\psi_{xy}$ and Lemma \ref{lem vertical}-\reff{lem vertical 3},
\begin{eqnarray}\label{u mono b}
u_x(x,-d)>0~~~~{\rm for~~} x<0,~~~~{\rm and} ~~~~ u_x(x,-d)<0~~~~{\rm for~~} x>0.
\end{eqnarray}
Then the fact that $U(-d)\geq 0$ and $U(0)>0$ together with the monotonicity properties \reff{u mono s} and \reff{u mono b} yields $u>0$ on the free surface and on the bottom. Differentiating \reff{equ psi} with respect to $y$ yields that
$u$ satisfies
\[
\triangle u+\gamma'(\psi)u=\gamma'(\psi) c\leq 0.
\]
Finally considering $U(y)\geq 0$ for $y\in[-d,0]$ and the properties (P\ref{u to 0})-(P\ref{u to U}), we can deduce $u>0$ in $\bar\Omega_\eta$ by using the strong maximum principle as in the proof of Lemma \ref{lem vertical}-\reff{lem vertical 2}, completing the proof.
\end{proof}
The above lemma considers the case $c\geq\sqrt{\lambda+2\Gamma(p_0)}$, that is the underlying flow is favorable. For the remaining cases, we can only derive some conclusions on the horizontal velocity $u$ for some special vorticity function $\gamma$ and for $\lambda$ sufficiently close to the corresponding $\lambda_c$ defined in \reff{def lambda c}. To be exact, we prove
\begin{lemma}\label{lem horizontal small}
Let $\lambda_c$ be determined in \reff{def lambda c} by the vorticity function $\gamma(p)$ with $p\in[0,\abs{p_0}]$.
\begin{enumerate}[(a)]
\item\label{ux} If $\gamma'(p),\gamma''(p)\leq0$ for $p\in[0,\abs{p_0}]$ and $\gamma(0)\sqrt{\lambda_c}>-g$, then there exists $\varepsilon_1>0$ such that for $\lambda\in(\lambda_c,\lambda_c+\varepsilon_1)$, we have
\begin{eqnarray*}
u_{x}> 0~~ {\rm for~~} (x,y)\in\Omega_-, ~~~{\rm and~~} u_{x}<0~~ {\rm for~~} (x,y)\in\Omega_+;
\end{eqnarray*}
\item\label{uy} If $\gamma(p)<0$ for all $p\in[0,\abs{p_0}]$, then there exists $\varepsilon_2>0$ such that for $\lambda\in(\lambda_c,\lambda_c+\varepsilon_2)$,
we have
\begin{eqnarray*}
u_{y}> 0~~ {\rm for~~} (x,y)\in\Omega.
\end{eqnarray*}
\end{enumerate}
\end{lemma}
\begin{proof}
$\reff{ux}$ Similar results have been obtained for periodic waves in \cite{Enonlinearity}, so we will adapt the proof there to our present solitary wave case.
We prove only the result for $(x,y)\in\Omega_+$ and the rest can be proved similarly. Note it suffices to prove that the stream function $\psi$ satisfies
\begin{eqnarray*}
\psi_{xy}<0~~ {\rm for~~} (x,y)\in\Omega_+.
\end{eqnarray*}
Since $\psi(x,\eta(x))=0$, we have $\psi_x=-\eta'\psi_y$ on the surface, which combined with \reff{Bernouli} gives
\[
\psi_y(x,\eta(x))=-\sqrt{(\lambda-2g\eta)/(1+\eta'^2)}.
\]
Differentiating the above equality and $\psi_x=-\eta'\psi_y$ along the surface gives respectively
\begin{eqnarray*}
\psi_{xy}+\eta'\psi_{yy}=-\partial_x\sqrt{\frac{\lambda-2g\eta}{1+\eta'^2}}\quad {\rm and}\quad\psi_{xx}+\eta'\psi_{yy}+2\eta'\psi_{xy}+\eta''\psi_y=0,~~~{\rm on ~~}y=\eta(x).
\end{eqnarray*}
Moreover on the surface one has
\[
\psi_{xx}+\psi_{yy}=-\gamma(0).
\]
Combination of the above three equalities gives
\[
\psi_{xy}(x,\eta(x))=\frac {\eta'\inner{g(1-\eta'^4)+2\eta''(\lambda-2g\eta)+\gamma(0)\sqrt{\lambda-2g\eta}(1+\eta'^2)^{3/2}}}
{\sqrt{\lambda-2g\eta}(1+\eta'^2)^{5/2}}.
\]
In view of \reff{eta small} and $\gamma(0)\sqrt{\lambda_c}>-g$, we have when $\lambda$ is sufficiently close to $\lambda_c$ that
\[
g(1-\eta'^4)+2\eta''(\lambda-2g\eta)+\gamma(0)\sqrt{\lambda-2g\eta}(1+\eta'^2)^{3/2}>0.
\]
This gives $\psi_{xy}<0$ on the surface since $\eta'<0$ for $x>0$ and $\sqrt{\lambda-2g\eta}(1+\eta'^2)^{5/2}>0$. Since $v(0,y)=0$, we have $\psi_{xy}=-v_y=0$ on the line $x=0$. On the bottom $\psi_{xy}<0$ holds due to Lemma \ref{lem vertical}-\reff{lem vertical 3}. And $\psi_{xy}=u_x\rightarrow 0$ as $x\rightarrow\infty$ since $u(x,y)\rightarrow U(y)$ as $x\rightarrow\infty$. Finally $\psi_{xy}$ satisfies
\begin{eqnarray*}
(\triangle+\gamma')\psi_{xy}=-\gamma''\psi_x\psi_y\geq 0.
\end{eqnarray*}
The conclusion follows from the maximum principle.
$\reff{uy}$ Recall from \cite{HurSolitary} that for $\lambda=\lambda_c+\varepsilon$, there exists a function $w^\varepsilon(q,p)$ such that $u(x,y)$ is determined by
\begin{eqnarray}\label{expression u}
u(x,y)=c-\frac{1}{(\lambda+2\Gamma(-\psi(x,y)))^{-1/2}+\varepsilon w^{\varepsilon}_p(\sqrt{\varepsilon}x,-\psi(x,y))},
\end{eqnarray}
where $w^{\varepsilon}_p(q,p)$ stands for the differentiation of the function $w^{\varepsilon}(q,p)$ with respect to $p$, and the family $\{w^{\varepsilon}(q,p); 0\leq \varepsilon<1\}$ satisfies that for $\varepsilon$ small
\begin{eqnarray}\label{property w}
\sup\Bigl\{\abs{\partial^j_q\partial^k_p w^{\varepsilon}(q,p)};j+k\leq2, (q,p)
\in\mathbb{R}\times[p_0,0] \Bigr\}\leq r,
\end{eqnarray}
with $r>0$ some constant independent of $\varepsilon$. Differentiating \reff{expression u} with respect to $y$ gives
\begin{eqnarray*}
u_y=\frac{-\psi_y\inner{-\gamma(\psi)\inner{\lambda+2\Gamma(-\psi)}^{-3/2}+\varepsilon w^\varepsilon_{pp}(\sqrt{\varepsilon}x,-\psi)}}
{\inner{\inner{\lambda+2\Gamma(-\psi)}^{-1/2}+\varepsilon w^\varepsilon_p(\sqrt{\varepsilon}x,-\psi)}^2}.
\end{eqnarray*}
Since $\gamma(p)<0$, if we denote $\gamma_{\max}=\max_{p\in[0,\abs{p_0}]}\gamma(p)$, then $\gamma_{\max}<0$ and $\Gamma_{\max}=\max_{p\in[p_0,0]}\Gamma(p)>0$. As a result, observing
\begin{eqnarray*}
-\gamma(\psi)\inner{\lambda+2\Gamma(-\psi)}^{-3/2}+\varepsilon w^\varepsilon_{pp}(\sqrt{\varepsilon}x,-\psi)\geq-\gamma_{\max}\inner{\lambda+2\Gamma_{\max}}^{-3/2}+\varepsilon w^\varepsilon_{pp}(\sqrt{\varepsilon}x,-\psi)
\end{eqnarray*}
and
\begin{eqnarray*}
\lim_{\varepsilon\rightarrow 0}\inner{-\gamma_{\max}\inner{\lambda+2\Gamma_{\max}}^{-3/2}+\varepsilon w^\varepsilon_{pp}(\sqrt{\varepsilon}x,-\psi)}
=-\gamma_{\max}\inner{\lambda_c+2\Gamma_{\max}}^{-3/2}
\end{eqnarray*}
due to \reff{property w},
we get, in view of $-\gamma_{\max}\inner{\lambda_c+2\Gamma_{\max}}^{-3/2}>0$, that
\begin{eqnarray*}
-\gamma(\psi)\inner{\lambda+2\Gamma(-\psi)}^{-3/2}+\varepsilon w^\varepsilon_{pp}(\sqrt{\varepsilon}x,-\psi)>0
\end{eqnarray*}
when $\varepsilon$ is sufficiently small. This combined with $-\psi_y>0$ gives $u_y>0$ for $\varepsilon$ sufficiently small.
\end{proof}
\begin{lemma}\label{u mono streamline}
If $\gamma(p)<0,\gamma'(p),\gamma''(p)\leq0$ for $p\in[0,\abs{p_0}]$ and $\gamma(0)\sqrt{\lambda_c}>-g$, then for $\lambda\in(\lambda_c,\lambda_c+\varepsilon_0)$ with $\varepsilon_0=\min\{\varepsilon_1,\varepsilon_2\}$, along every streamline $y=\sigma_p(x)$ with $p\in(p_0,0)$, the horizontal velocity $u$ is strictly decreasing in $\Omega_+$ and strictly increasing in $\Omega_-$ as a function of $x$.
\end{lemma}
\begin{proof}
Since
\begin{eqnarray*}
\frac{d }{dx}u(x,\sigma_p(x))=u_x+u_y\sigma_p'(x)=u_x+u_y\frac{v}{u-c}
\end{eqnarray*}
due to \reff{sigma}, the conclusion follows immediately from Lemma \ref{lem vertical} and Lemma \ref{lem horizontal small}.
\end{proof}
\section{Main result}\label{sec r}
Recalling that $y=l(y_0)$ is the streamline asymptote introduced at the end of Section \ref{sec p}, we are now ready to state our main result describing the particle
trajectories in a solitary wave with negative vorticity.
\begin{theorem}\label{thm}
Let the flux $p_0<0$, the vorticity function $\gamma\in C^2(0,\abs{p_0})$
be given, and let $\lambda_c$ be determined as in \reff{def lambda c}. Assume that $\gamma(p)<0,\gamma'(p)\leq0$ for all $p\in [0,\abs{p_0}]$, and that $\lambda>\lambda_c$ such that nontrivial solitary wave solutions to \reff{psi y negative}-\reff{psi x to 0} exist. Then the following results hold.
\begin{enumerate}[(a)]
\item\label{under crest} Any particle above the bed reaches at some instant $t_0$ the location $(X_0,Y_0)$ below the wave crest $(X_0,\eta(0))$;
\item\label{Tra positive} For $c\geq\sqrt{\lambda+2\Gamma(p_0)}$, as time $t$ runs on $(-\infty,t_0)$, the particle above the flat bed moves to the right and upwards, while for $t>t_0$ the particle moves to the right and downwards, as in Figure \ref{fig1}-(a); the particle on the flat bed moves in a straight line to the right at a positive speed;
\item\label{Tra nonpositive} If assume additionally that $\gamma(0)\sqrt{\lambda_c}>-g$ and $\gamma''(p)\leq 0$ for all $p\in [0,\abs{p_0}]$, then there exists $\varepsilon_0>0$ such that for $\lambda\in(\lambda_c,\lambda_c+\varepsilon_0)$ and
\begin{enumerate}[(i)]
\item\label{Tra mixed} for $\sqrt{\lambda}<c<\sqrt{\lambda+2\Gamma(p_0)}$, there does not exist a single pattern for all the particles above the flat bed; if $u(x,-d)\geq 0$, depending on the relation between the asymptote $y=l(Y_0)$ and the zero point $y_*$ of $U(y)$, some particles move to the right along a path with a single hump as described in \reff{Tra positive}, some move along a single-loop path to the left, as in Figure \ref{fig1}-(b); if $u(x,-d)<0$, there are additionally some particles moving to the left along a single-hump path; see Figure \ref{fig1}-(c);
\item\label{Tra negative} for $c\leq\sqrt{\lambda}$, depending on the signs of $u(0,\eta(0))$ and $u(0,-d)$, there are three possibilities for the particles above the flat bed; see Figure \ref{fig3};
\item\label{Tra bottom} for a particle on the flat bed in both the cases (i) and (ii), if $u(x,-d)\leq0$ it moves to the left in a straight line, while if $u(x,-d)>0$ it firstly has a backward-forward pattern of motion and then moves to the left in a straight line;
\end{enumerate}
\item\label{Tra infinity} The particle trajectory is strictly above the asymptote $Y=l(Y_0)$ of the streamline $Y=\sigma_p(X-ct)$ with $p=-\psi(0,Y_0)$.
\end{enumerate}
\end{theorem}
\begin{proof}
The path (past and future) $(X(t),Y(t))$ of a particle with location $(X(0),Y(0))$ at time $t=0$ is given by the solution of the non-autonomous system
\begin{eqnarray*}
\left\{
\begin{array}{ll}
\dot X=u(X-ct,Y),\\
\dot Y=v(X-ct,Y).
\end{array}
\right.
\end{eqnarray*}
Working in the moving frame $x=X-ct$ and $y=Y$, we transform the above system into
\begin{eqnarray}\label{equ x y}
\left\{
\begin{array}{ll}
\dot x=u(x,y)-c,\\
\dot y=v(x,y).
\end{array}
\right.
\end{eqnarray}
This is a Hamiltonian system with Hamiltonian $\psi(x,y)$ in view of \reff{def psi}, meaning that the solutions $(x(t),y(t))$ of \reff{equ x y} lie on the streamlines. All solutions of \reff{equ x y} are defined globally in time since the boundedness of the right-hand side prevents blow-up in finite time.
\reff{under crest} Since in the moving frame the wave crest is assumed to be located at $x=0$ in our setting, the sign of $x(t)$ describes the position of the particle with respect to the wave crest at time $t$: the particle is exactly below the crest if $x(t)=0$, is ahead of the crest if $x(t)>0$ while is behind the crest if $x(t)<0$. Note that $\dot x=u(x,y)-c\leq -\delta< 0$ throughout $\bar\Omega_{\eta}$. This uniform upper bound on $\dot x$ implies that $x(t)\rightarrow \mp\infty$ as $t\rightarrow \pm\infty$ and there is a unique time $t_0$ such that $x(t_0)=0$. That is, to each fluid particle moving within the water there corresponds a unique time $t_0\in\mathbb{R}$ so that at $t=t_0$ the particle is exactly below the wave crest, while afterwards it is located behind the wave crest, the wave crest being behind the particle for $t<t_0$.
In the subsequent of the proof, we always assume that the particle is located below the crest at time $t=t_0$ at the location $(X_0,Y_0)=(X(t_0),Y(t_0))$, or equivalently $(x(t_0),y(t_0))=(0,Y_0)$ which implies in fact $X_0=ct_0$.
\reff{Tra positive}
Assume $c\geq\sqrt{\lambda+2\Gamma(p_0)}$. For a particle located above the flat bed, we prove only the statement for $t>t_0$ and the case $t<t_0$ can be proved similarly. Since $x(t)$ is strictly decreasing, one has $X(t)-ct=x(t)<0$ for $t>t_0$, which combined with Lemma \ref{lem vertical}-\reff{lem vertical 2} implies $\dot Y<0$ for $t>t_0$. Furthermore we have $\dot X>0$ for all time by Lemma \ref{lem horizontal}. That means the particle moves to the right and downwards as time runs on $(t_0,+\infty)$. The statement for particles on the bed follows from Lemma \ref{lem vertical}-\reff{v=0}
and Lemma \ref{lem horizontal}. The particle path for this case is depicted in Figure \ref{fig1}-(a).
\begin{figure}
\begin{center}
\includegraphics{picture.1}\qua
\includegraphics{picture.2}\quad
\includegraphics{picture.3}
\end{center}
\caption{\small {Particle path with a single: (a) hump to the right; (b) loop to the left; (c) hump to the left.}}\label{fig1}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics{picture.4}\hspace{1.5cm}
\includegraphics{picture.5}
\end{center}
\caption{\small {Particle path above the flat bed in a small solitary wave with a mixed underlying current.}}\label{fig2}
\end{figure}
\reff{Tra nonpositive}-(i) Since
$\sqrt{\lambda}<c<\sqrt{\lambda+2\Gamma(p_0)}$, we have $U(0)>0$
and $U(-d)<0$, thus $U(y)$ has a unique zero point, say at
$y_*$.
Assume first that $u(0,-d)\geq 0$. The monotonicity properties of $u$ provided by Lemma \ref{lem horizontal small}-\reff{ux} and \reff{u mono b} show that in $\bar\Omega_+$ the level set $\{u=0\}$ consists of a continuous curve ${\mathcal C}_+$ in the moving frame. The curve is confined between $y=-d$ and $y=y_*$, and can be parameterized by $x=h(y)$ with $h'(y)>0$, $h(y)\rightarrow +\infty$ as $y\rightarrow y_*$, and $h(-d)\geq 0$ with the equality holding when $u(x,-d)=0$. In $\bar\Omega_-$ the level set $\{u=0\}$ is given by the reflection ${\mathcal C}_-$ of the curve ${\mathcal C}_+$ across the line $x=0$. Below the curve ${\mathcal C}_+$ and ${\mathcal C}_-$ we have $u<0$ (including the bottom), while above and between the two curves we have $u>0$. Furthermore, in view of Lemma \ref{u mono streamline}, if a streamline and the curve ${\mathcal C}_+$ (rep. ${\mathcal C}_-$) intersect, they intersect exactly once.
Recall that $(x(t),y(t))$ lies on the streamline $y=\sigma_p(x)$
with $p=-\psi(0,Y_0)$ since $(x(t_0),y(t_0))=(0,Y_0)$, and recall
that $y=l(Y_0)$ is the asymptote of the streamline passing through the point $(0,Y_0)$.
In virtue of Corollary \ref{cor streamline}, the path $(x(t),y(t))$ is located below $y=Y_0$ and above the asymptote $y=l(Y_0)$ for all time $t$. If $l(Y_0)\geq y_*$, then we have from the above arguments that $u>0$ for all time $t$, while $v<0$ if $t>t_0$ and $v>0$ if $t<t_0$. This is the same situation as described in \reff{Tra positive}, so that the particle moves to the right along a path with a single hump, as depicted in Figure \ref{fig1}-(a).
If $l(Y_0)<y_*$, the particle trajectories above the flat bed are as shown in Figure \ref{fig1}-(b). Indeed, in this case, as time $t$ increases from $-\infty$, the path $(x(t),y(t))$ intersects successively the curve ${\mathcal C}_+$ from below at $t=t_+$, the vertical line $x=0$ from right at $t=t_0$, and the curve ${\mathcal C}_-$ from above at $t=t_-$. In the time interval $t\in(-\infty,t_+)\cup(t_-,+\infty)$ we know that $u<0$ so that in the physical variables $(X,Y)$ the particle $(X(t),Y(t))$ moves to the left. In the time interval $t\in(t_+,t_-)$ we have $u>0$ so that $(X(t),Y(t))$ moves to the right. Also we have $v>0$ when $t<t_0$ so that $(X(t),Y(t))$ moves up, while when $t>t_0$ we have $v<0$ so that $(X(t),Y(t))$ moves down. Thus in this case the particle above the flat bed moves to the left along a path with a single loop.
It remains to treat the case $u(0,-d)<0$. Note $u(0,\eta(0))>0$ in virtue of $U(0)>0$ and \reff{u mono s}. Thus, in view of Lemma
\ref{lem horizontal small}-\reff{uy}, there exists a unique
$\tilde y\in(-d, \eta(0))$ such that $u(0,\tilde y)=0$. Observe
$\tilde y<y_*$ since $u_x<0$ in $\Omega_+$ and $u(x,y)\rightarrow U(y)$ as $x\rightarrow \infty$. The corresponding
curve ${\mathcal C}_+$ is now located between $y=y_*$ and
$y=\tilde y$, intersecting the line $x=0$ at $y=\tilde y$. For
$Y_0>\tilde y$, the paths $(X(t),Y(t))$ are similar as encountered
in the case when $u(0,-d)\geq 0$. While for $Y_0\leq \tilde y$, we
have $(x(t),y(t))$ is below ${\mathcal C}_+$ and ${\mathcal C}_-$,
so that $u<0$ for all the time. As a result, the particles move to
the left along a single-hump path; see Figure \ref{fig1}-(c).
We depict all the possible trajectories in this case in Figure \ref{fig2}.
\reff{Tra nonpositive}-(ii) This case can be treated similarly as that in the
above, so we omit the details and show the particle trajectories in
Figure \ref{fig3}.
\begin{figure}
\begin{center}
\includegraphics{picture.6}\quad\quad
\includegraphics{picture.7}\quad \quad
\includegraphics{picture.8}
\end{center}
\caption{\small {Particle path above the flat bed in a small solitary wave with an adverse underlying current.}}\label{fig3}
\end{figure}
\reff{Tra nonpositive}-(iii) The conclusions for particles on the flat bed
follow from Lemma \ref{lem vertical}-\reff{v=0} and \reff{u mono b}.
\reff{Tra infinity} Observing $\dot Y=v(X-ct,Y)$, $X(t)-ct=x(t)\rightarrow -\infty$ as $t\rightarrow+\infty$ and $v(x,y)$ converges to $0$ as $\abs{x}\rightarrow \infty$ uniformly in $y$, we have the existence of some $\alpha$ such that $\lim_{t\rightarrow+\infty}Y(t)=\alpha$. Since $\psi(x,y)$ is the Hamiltonian function of the system \reff{equ x y}, we have
\[
\psi(x(t),y(t))=\psi(x(t_0),y(t_0))=\psi(0,Y_0).
\]
Thus $y(t)=\sigma_p(x(t))$ with $p=-\psi(0,Y_0)$. Consequently
\[
\lim_{t\rightarrow +\infty}Y(t)=\lim_{t\rightarrow +\infty}y(t)=\lim_{t\rightarrow +\infty}\sigma_p(x(t))=l(Y_0).
\]
The proof is completed.
\end{proof}
\noindent{\bf Remark}
For the path with a single loop, if we define the size of the loop by the net horizontal distance moved by the particle between the two instants $t_+$ and $t_-$ when its horizontal velocity changes sign, that is,
\begin{eqnarray*}
X(t_-)-X(t_+),
\end{eqnarray*}
then the size decreases with depth. In fact it can be computed by
\begin{eqnarray*}
X(t_-)-X(t_+)=\int_{t_+}^{t_-}\frac{d X}{dt}dt=\int_{t_+}^{t_-}u\inner{x,\sigma_p(x)}dt.
\end{eqnarray*}
Since $u_y>0$ and $\frac{d \sigma_p(x)}{dp}=-\frac{1}{\psi_y}>0$, we have $X(t_-)-X(t_+)$ decreases as $p$ decreases.
\section{Examples}\label{sec e} In this final section we give two examples of the vorticity function which satisfy the conditions imposed in our main theorem.
\subsection{Negative constant vorticity} In the case of negative constant vorticity $\gamma(p)=-\omega_0$ for $p\in[0,\abs{p_0}]$ with $\omega_0>0$, we have obviously $\gamma'(p),\gamma''(p)\leq 0$. So it remains to verify $\gamma(0)\sqrt{\lambda_c}>-g$, or equivalently, $\lambda_c<g^2/\omega_0^2$. Recall from \reff{def lambda c} that
$\lambda_c>-2\Gamma_{\min}$ is such that
\begin{eqnarray*}
\int_{p_0}^0\frac{dp}{\inner{\lambda_c+2\Gamma(p)}^{3/2}}=\frac{1}{g}
\end{eqnarray*}
holds, where $\Gamma(p)=-\omega_0 p$ and $\Gamma_{\min}=0$ in this case by \reff{def Gamma}. Set
\[
F(\lambda)=\int_{p_0}^0\frac{dp}{{\inner{\lambda-2\omega_0 p}}^{3/2}}.
\]
Then direct computation shows that $F'(\lambda)<0$, $F(\lambda)\rightarrow +\infty$ as $\lambda\rightarrow 0+$, and
\begin{eqnarray*}
\lim_{\lambda\rightarrow g^2/\omega_0^2}F(\lambda)=\frac{1}{g}-\frac{1} {\sqrt{g^2-2\omega_0^3p_0}}<\frac{1}{g}.
\end{eqnarray*}
Thus there exists a unique $\lambda_c\in(0,g^2/\omega_0^2)$ such that $F(\lambda_c)=1/g$.
\subsection{Negative affine linear vorticity} For the vorticity $\gamma(p)=-a p+b$ with $a>0$ and $b<0$,
we have obviously $\gamma(p)<0$, $\gamma'(p),\gamma''(p)\leq 0$ for $p\in[0,\abs{p_0}]$ and
\[
F(\lambda)=\int_{p_0}^0\frac{dp}{{\inner{\lambda+2\Gamma(p)}}^{3/2}}=\int_{p_0}^0\frac{dp}{{\inner{\lambda+ap^2+2b p}}^{3/2}},
\]
with $\lambda>-2\Gamma_{\min}=0$. Then we may verify as above that $F(\lambda)$ is strictly decreasing with $\lim_{\lambda\rightarrow 0}F(\lambda)=+\infty$ and $\lim_{\lambda\rightarrow g^2/b^2}F(\lambda)<0$. This gives the existence of $\lambda_c\in(0,g^2/b^2)$ such that $F(\lambda_c)=1/g$ and consequently $\gamma(0)\sqrt{\lambda_c}>-g$.
\subsection*{Acknowledgements} This work was supported in part by
Foundation of WUST(2010xz019) and by NSF grants of China 10901126.
|
\section{Introduction}\label{sect:intro}
Solar flares are believed to be powered by magnetic reconnection high in the corona, which accelerate particles.
Particles, and in particular electrons, travel down field lines and emit bremsstrahlung hard X-rays (HXR) as they penetrate the denser chromosphere.
Hence, the HXR radiation from the footpoints contain much information about the accelerated electrons (though convoluted with transport effects), such as energy content
\cite{Brown1971,Holman2003,PSH2005}.
The first observations of HXR footpoints were made by \inlinecite{Hoyng1981} on SMM.
Yohkoh HXT later characterized the ``standard'' flare model: two footpoints and an above-the-looptop source \cite{Masuda1994}.
The above-the-looptop source is rarely observed, although that could be due to observational constraints.
Statistical studies of HXR footpoints (and looptop) sources' spectral indices using the Yohkoh satellite \cite{Sakao1994,Petrosian2002} have been carried out in the past.
Some of their results seem to indicate that the spectral indices of two neighbouring footpoints could differ by as much as 1 or even 2!
Such differences cannot be explained easily by transport mechanisms (see {\it e.g.} Appendix \ref{appendix:ColumnDensity}).
Yohkoh HXT's results were compromised by the fact that it had only had 4 energy channels, and had to deal with issues like thermal contamination of the lowest channel(s), and the sometime poor statistics of the upper ones.
The Ramaty High Energy Solar Spectroscopic Imager (RHESSI, \opencite{Lin2002}) offers an unprecedented combination of spectral resolution (1 keV in the 3--100 keV range), spatial resolution (2.3$''$), temporal evolution ($\approx$2s), and sensitivity comparable to Yohkoh HXT's.
Previous RHESSI imaging spectroscopy results include: \inlinecite{Krucker2002}, \inlinecite{Emslie2003}, \inlinecite{Marina2006}.
A similar time variation of spectra in footpoints is observed, although in some cases, the spectrum in one footpoint is steeper than in the other one (by about 0.3 in spectral index, in \inlinecite{Emslie2003}.
\inlinecite{Emslie2003} suggest that the discrepancy could be due to a difference in column densities of the electron population, as they propagate down an asymmetric loop (cf Appendix \ref{appendix:ColumnDensity}).
All RHESSI papers so far discuss only a single event, or a few events, but no statistical study of the footpoints has been done so far, that exploit RHESSI's large database of observed flares.
These results should help constrain energy release and particle acceleration in the flare model.
\section{Event Selection}\label{sect:selection}
\begin{figure}
\centering
\includegraphics[width=11cm]{exampleflare.ps}
\caption{RHESSI time profile (top), 30--33 keV image (bottom left) and footpoint spectra (Bottom right) of the 13 July 2005 M2.7 flare.
The gray bars represent the time intervals chosen for our analysis: a ``Peak Times'' 16-second interval around 14:14:10, and two ``Whole Peaks''
from 14:13:47 to 14:19:50, and from 14:21:22 to 14:24:41.
The third HXR peak ($\approx$14:30) was discarded as it was faint, and there was an attenuator state change (Lin \etal, 2002) during it, further complicating the analysis.
The sun was eclipsed by the Earth until $\approx$14:13.}
\label{fig:exampleflare}
\end{figure}
RHESSI has been in orbit for six years now.
Since its launch on 5 February 2002, it has observed more than 20000 flares.
For this study, the strongest flares will be taken into considerations, as they have better count statistics.
More specifically, flares which show substantial emission above 50 keV (where solar flare thermal components are always negligible), enough to produce images of good enough quality to be used in imaging spectroscopy (2000 counts is about the minimum for a reliable single image reconstruction with two sources).
To achieve our objectives, we will use the simplest possible events, those that show only two footpoints.
The HESSI Experimental Data Center (HEDC, \opencite{PSH2002}) was used to find our events.
It was queried for all flares between 13 February 2002 and 1 July 2006 which had 2 or more sources, and with peak GOES flux above M1.0 level.
More than 1100 flares corresponded to that description.
Each were individually examined, in particular the time vs. energy panels of RHESSI images that HEDC automatically produces for each flare (one minute accumulations over whole flare duration).
The flares that were retained were those that visually displayed two footpoints in HEDC images above 50 keV.
53 flares were kept (see Table \ref{table:1}).
\begin{table}
\centering
\begin{tiny}
\caption{Flares studied, GOES X-ray class, and position on the sun in both arcseconds from suncenter and heliographic coordinates.}
\label{table:1}
\begin{tabular}{c c c c c c}
\hline
Flare & Date and& GOES & \multicolumn{3}{c}{Location on Sun} \\
number& time & class & X [$''$] & Y[$''$] & Lat/Lon \\
\hline
1 & 2002/02/20 11:06:12 & C7.5 & 904.4 & 261.5 & N13 W73 \\
2 & 2002/03/17 10:15:36 & M1.4 & -342.2 & -239.2 & S20 E22 \\
3 & 2002/03/17 19:28:44 & M4.4 & -264.8 & -232.9 & S20 E17 \\
4 & 2002/04/10 12:28:04 & M8.8 & -20.8 & 434.5 & N20 E01 \\
5 & 2002/04/10 19:02:48 & M1.8 & -348.2 & 377.5 & N17 E22 \\
6 & 2002/05/31 00:07:08 & M2.4 & -817.5 & -475.4 & S30 E87 \\
7 & 2002/06/01 03:53:40 & M1.6 & -414.8 & -293.0 & S18 E27 \\
8 & 2002/06/02 11:44:32 & M1.0 & -148.8 & -300.3 & S18 E09 \\
9 & 2002/07/17 07:02:48 & M9.2 & 288.5 & 246.5 & N19 W18 \\
10 & 2002/07/18 03:32:56 & M2.5 & 421.4 & 264.5 & N20 W28 \\
11 & 2002/07/23 00:28:04 & X5.1 & -868.1 & -235.3 & S12 E70 \\
12 & 2002/07/29 10:39:08 & M5.1 & 238.9 & -291.3 & S12 W14 \\
13 & 2002/07/31 01:48:40 & M1.4 & 558.4 & -220.2 & S08 W36 \\
14 & 2002/08/03 19:04:36 & X1.2 & 899.8 & -265.8 & S15 W80 \\
15 & 2002/08/21 01:39:16 & M1.6 & 689.2 & -246.7 & S10 W47 \\
16 & 2002/08/22 01:52:00 & M5.9 & 798.4 & -266.5 & S12 W59 \\
17 & 2002/09/08 01:39:08 & M1.6 & -908.8 & -193.1 & S09 E75 \\
18 & 2002/09/27 03:34:28 & M1.0 & -694.7 & 142.9 & N13 E47 \\
19 & 2002/12/04 22:47:00 & M2.5 & -836.7 & 227.7 & N13 E61 \\
20 & 2003/04/23 01:01:56 & M5.1 & 261.8 & 366.8 & N17 W16 \\
21 & 2003/05/29 01:04:40 & X1.2 & 494.4 & -106.6 & S07 W31 \\
22 & 2003/06/17 22:53:40 & M6.8 & -790.4 & -138.6 & S07 E57 \\
23 & 2003/07/17 08:19:46 & C9.7 & -206.9 & 178.2 & N15 E13 \\
24 & 2003/10/23 08:47:20 & X5.4 & -904.8 & -317.4 & S18 E81 \\
25 & 2003/10/24 02:48:32 & M7.7 & -865.1 & -341.1 & S19 E71 \\
26 & 2003/10/29 20:43:20 & X10 & 90.8 & -381.3 & S18 W05 \\
27 & 2003/11/01 22:33:04 & M3.3 & 818.6 & -253.8 & S13 W60 \\
28 & 2003/11/03 09:49:16 & X3.9 & 917.8 & 130.3 & N08 W73 \\
29 & 2003/11/04 19:33:56 & M9.1 & 900.8 & -335.0 & S19 W81 \\
30 & 2004/01/06 06:22:32 & M2.7 & -972.6 & 88.3 & N05 E89 \\
31 & 2004/01/07 10:22:12 & M3.7 & -930.2 & 117.0 & N05 E73 \\
32 & 2004/04/06 13:22:48 & M2.3 & -261.5 & -170.6 & S16 E16 \\
33 & 2004/07/13 00:15:26 & M6.8 & 654.7 & 181.6 & N14 W45 \\
34 & 2004/07/23 21:19:28 & M1.8 & 125.5 & 6.6 & N05 W07 \\
35 & 2004/09/12 00:33:44 & M4.8 & -706.2 & -35.9 & N02 E47 \\
36 & 2004/10/30 03:30:09 & M3.5 & 316.8 & 145.2 & N12 W19 \\
37 & 2004/10/30 16:24:26 & M6.0 & 427.6 & 139.5 & N12 W26 \\
38 & 2004/10/31 05:32:03 & M2.4 & 540.6 & 152.6 & N12 W34 \\
39 & 2004/11/03 03:30:52 & M1.6 & -674.1 & 94.7 & N08 E44 \\
40 & 2004/11/06 00:30:48 & M9.5 & -79.3 & 83.4 & N08 E04 \\
41 & 2004/11/06 01:42:34 & M3.7 & -27.4 & 68.1 & N07 E01 \\
42 & 2004/11/10 02:09:44 & X2.6 & 700.1 & 91.4 & N07 W46 \\
43 & 2004/12/01 07:10:16 & M1.2 & -335.8 & 128.5 & N08 E20 \\
44 & 2005/01/15 06:28:31 & M8.4 & -106.1 & 295.5 & N12 E06 \\
45 & 2005/01/15 22:48:24 & X2.7 & 103.1 & 306.7 & N13 W06 \\
46 & 2005/01/17 10:00:23 & X3.9 & 430.0 & 292.3 & N13 W26 \\
47 & 2005/01/19 08:12:40 & M8.7 & 708.9 & 283.6 & N13 W48 \\
48 & 2005/01/19 10:21:08 & M2.5 & 679.6 & 339.9 & N16 W46 \\
49 & 2005/01/20 06:44:44 & X7.1 & 818.5 & 256.0 & N12 W59 \\
50 & 2005/07/13 14:14:14 & M2.7 & 909.4 & 168.9 & N11 W78 \\
51 & 2005/08/22 01:11:54 & M2.3 & 717.8 & -248.7 & S10 W50 \\
52 & 2005/08/22 17:07:34 & M5.2 & 801.4 & -241.6 & S11 W59 \\
53 & 2005/08/23 14:46:21 & M2.7 & 883.8 & -219.0 & S11 W71 \\
\hline
\end{tabular}
\end{tiny}
\end{table}
\section{Method of Analysis}\label{sect:method}
Imaging spectroscopy using CLEAN \cite{Hurford2002} and the OSPEX spectral analysis software were employed.
Imaging was done using collimators 3 to 8, yielding a formal image resolution of 7$''$ FWHM.
The time intervals and energy intervals were chosen as follow.
For each flare, two types of time intervals were used:
\begin{itemize}
\item {\it ``Peak flux'' time intervals:}
These are four RHESSI spin periods long (each RHESSI spin period being about 4s long), centered at the time of peak HXR (above 50 keV) flux.
The later is found using RHESSI Observing Summary data \cite{Schwartz2002}.
The peak flux time interval was taken to be this time of peak flux plus or minus two RHESSI spin periods (which are $\approx$4s long).
Of course, there can be only one such peak flux time interval per flare, resulting in 53 such peaks in our study.
Taking the flare of 13 July 2005 as an example (Figure \ref{fig:exampleflare}), the time interval of accumulation was about 14:14:05 to 14:14:22.
\item {\it ``Whole peak'' time intervals:}
Strong non-thermal peaks appearing in RHESSI spectrograms (or dynamic spectra) were selected over their whole {\it time interval},
defined as the time the HXR flux ($>$50 keV) is greater than 50\% of its peak value.
There can be many such ``Whole peak'' time intervals within the same flare.
And of course, for each flare, one of the ``Whole peak'' time interval envelops that flare's ``Peak flux'' time interval.
Again taking the flare of 13 July 2005 as an example (Figure \ref{fig:exampleflare}), the time intervals of accumulation were about
14:13:47 to 14:19:50, and 14:21:22 to 14:24:41.
In a few cases, time intervals did not contain {\bf two} footpoints (but only one, or sometimes three or more), and were hence discarded from the study.
\end{itemize}
For each time interval, the energy binning was chosen using the following semi-empirical approach:
\begin{itemize}
\item The {\it start (lowest)} energy was visually chosen by inspection of the RHESSI spectrogram: it is taken to be the point where the non-thermal emission starts to be clearly stronger than the thermal component.
\item The energy binning was taken to be pseudo-logarithmic, which each energy bin having at least 2000 counts above background, and the bin width being between 5\% and 20\% of the bin value.
This was crudely approximated using Observing Summary 4 second data rates.
The {\it end (highest)} energy was taken to be when the next energy bin could not achieve 2000 counts above background.
We had 4 to 18 (typically 10) energy bins to fit and obtain spectral indices and fluxes with.
\end{itemize}
Finally, for the results presented and discussed in this paper, only fittings deemed ``most reliable'' were kept.
``Most reliable'' meaning those which had at least 6 or more energy bins for footpoint spectral fitting, and a with a best-fit $\chi^2$ value of 5 or less.
Ultimately, 33 ``Peak Time'' and 89 ``Whole Peaks'' events were used to produce the results that we analyse here.
To limit {\it pulse pile-up} issues \cite{Smith2002}, care has been taken to discard times with high count rates ({\it i.e.} just before shutters moving in).
Moreover, as spectral fitting was usually done above 25 keV, only times with very strong emission (during which both attenuators are ``in'', or ``A3'' state) can potentially produce an additional component around 35 keV
(with only the thin shutter in (``A1'' state), detector countrates peak around 12 keV, and these can be pile-up to $\approx$24 keV.
In A3 state, the peak of the response is around $\approx$18 keV counts. These photons can pile up and appear as $\approx$36 keV photons).
In 20 of our events were the contribution of pile-up photons in certain energy channels larger than 15\%.
As pile-up typically makes two thermal photons appear as a single higher energy photon, imaging piled-up photons would place them at the location of the thermal source.
In 19 of these 20 cases, the thermal source was spatially distinct from the HXR footpoints, thereby little influencing our results.
In the remaining case, the thermal source overlapped with the non-thermal HXR footpoints (within our 7$''$ spatial resolution), and spectral fitting was done above 40 keV to eliminate any contamination by piled-up low-energy photons.
Table \ref{table:2} is a list of all parameters obtained for each of our events.
Subscripts 1 and 2 refer to the value of the leading and trailing footpoints, respectively (as determined by their heliographic longitude).
In a few cases, the subscripts {\it strong} and {\it weak} were also used.
They refer to the value of the strongest and weakest footpoints, respectively (as determined by their 50-keV flux).
Presenting all possible combination of scatter plots is prohibitive (they can be all found at a website\footnote{\url{http://sprg.ssl.berkeley.edu/~shilaire/FootPointProject/htmlsummaries/browser.html}}).
Only the most relevant have been presented, but all have of course been examined, and an exhaustive table of computed correlation coefficients can be found in Section \ref{sect:obs:corr}.
\begin{table}
\centering
\caption{Measured and derived quantities.}
\label{table:2}
\begin{tabular}{c l}
\hline
Symbol & Name or decription \\
\hline
$\gamma_1$, $\gamma_2$ & Spectral indices of both footpoints, as obtained by fitting \\
& a power-law using the OSPEX from the {\it Solarsoft} suite of\\
& routines \\
$\Delta\gamma$ & Spectral index difference $\Delta\gamma={\gamma}_1-{\gamma}_2$ between \\
& footpoints \\
$\bar{\gamma}$ & =$\frac{1}{2} ({\gamma}_1+{\gamma}_2)$: average spectral index \\
F$_{50,1}$, F$_{50,2}$ & 50-keV photon flux in both footpoints [ph/s/cm$^2$/keV]\\
F$_{50,\mathrm{tot}}$ & =F$_{50,1}$+F$_{50,2}$, total flux \\
F$_{50,\mathrm{r}}$ & =$\frac{\mathrm{F}_{50,1}}{\mathrm{F}_{50,2}}$: 50-kev flux ratio between footpoints\\
\hline
$dt$ & Duration or accumulation time [s]\\
GOES & GOES X-ray class, or flux [W m$^{-2}$] in the low 1--8{\AA} channel\\
\hline
Lat$_1$, Lat$_2$ & Heliographic longitude [degrees] of both footpoints. \\
& The footpoint with the largest longitude, (or ``leading'') is \\
& labelled ``1'', the other one ``2'' \\
Lon$_1$, Lon$_2$ & Heliographic latitude of both footpoints [degrees] \\
$s$ & Spherical separation between footpoints [Mm] \\
$\alpha$ & Angle between footpoints and solar equator \\
\hline
\end{tabular}
\end{table}
\section{Observations \& Discussion}\label{sect:obs}
\subsection{Spatial information}\label{sect:obs:spatial}
\begin{figure}
\centering
\includegraphics[width=10cm]{FlareDistributionOnSun.ps}
\caption{Distribution of our sample of 53 flares on the Sun.}
\label{fig:FlareDistribution}
\end{figure}
{\bf Flare distribution on the sun:} Figure \ref{fig:FlareDistribution} shows the spatial distribution of our 53 flares on the Sun.
As already known, flares occur predominently at $\pm$15 degrees of latitude, and there is no marked longitudinal dependence.
The slight lack of events at high longitudes is very probably due to observational bias:
with our imaging method (CLEAN with detectors 3 and above), flares with footpoint separation smaller than $\approx$10$''$ appear to be single-footpoint flare, and are not selected.
Projection effects near the solar limb reduces the apparent footpoint separation, causing some of these flares to be discarded.
\subsection{Spectral Information}\label{sect:obs:spectral}
\begin{figure}
\centering
\includegraphics[width=10cm]{HistoGamma.ps}
\caption{Histogram of spectral indices $\gamma$. The gray bars are for ``Whole Peaks'' events, the black ones for ``Peak Times'' events.
The bin size (0.2) was taken to be larger than the average error of 0.16.}
\label{fig:HistoGamma}
\end{figure}
Figure \ref{fig:HistoGamma} shows that spectral indices are generally harder for ``Peak Times'' events, which is of course no surprise,
as it is a natural consequence of the {\it Soft-Hard-Soft} behaviour observed in a majority of flares,
where the flattest spectral index corresponds to the time of most intense HXR emission (see {\it e.g.} \inlinecite{Paolo2004}).
Flares are very seldom harder than $\gamma\approx$2.4 in photon spectral index (see {\it e.g.} \inlinecite{Kasparova2005} and references therein).
The distribution at high $\gamma$ in Figure \ref{fig:HistoGamma} is not to be trusted, as it is distorted by observational bias:
only flares with sufficient HXR emission above 50 keV , {\it i.e.} flares with hard spectra, were used in our study.
Another observational constraint is the instrument's dynamic range $DR$:
The weakest footpoint is visible if F$_{\mathrm{weak}} \geq \frac{\mathrm{F}_{\mathrm{strong}}}{DR}$.
Using a conservative dynamic range of $\approx$5 for RHESSI, it means that if a footpoint is weaker than the other one by a factor 5 or more, it will not be imaged.
\begin{figure}
\centering
$\begin{array}{c@{\hspace{1in}}c}
\includegraphics[width=9.3cm]{gVsF.ps} \\
\includegraphics[width=9.3cm]{DgVsFr.ps} \\
\end{array}$
\caption{Top: $\bar{\gamma}$ vs. F$_{50,tot}$, with 1-$\sigma$ error bars.
Bottom: $\Delta\gamma$ and $|\Delta\gamma|$ vs. F$_{50,r}$, with 1-$\sigma$ error bars (see Table \ref{table:2} for an explanation of all quantities).
The data points in gray are the ones for which $|\Delta\gamma|<\sigma_{\Delta\gamma}$.
The thick, gray line is a linear regression to the $\Delta\gamma$ vs. F$_{50,\mathrm{r}}$ ``Peak times'' data.}
\label{fig:main}
\end{figure}
Figure \ref{fig:main} displays scatter plots of the average spectral indices ($\bar{\gamma}$) or spectral index differences ($\Delta\gamma$)
versus the total 50-keV flux (F$_{50,\mathrm{tot}}$) or the 50-keV flux ratio (F$_{50,\mathrm{r}}$) of our events, with error bars.
And indeed, no event shows a flux ratio greater than 5 (or smaller than 0.2).
Very few footpoint pairs have spectral index difference greater than 0.6, and none above 0.8.
This fact could not be attributed to observational effects.
During our data reduction, a few events with $\Delta\gamma$ larger than 1 were found, but they were discarded because of poor statistics (large $\chi^2$ fitting parameter) and/or the appearance of a third source.
\begin{figure}
\centering
\includegraphics[width=10cm]{HistoDg3.ps}
\caption{Histogram of spectral index differences $\Delta\gamma$.
Crosshatched: ``Peak Time'' events. Solid black: ``Whole Peak'' events. Solid gray: ``Whole Peak'' events (only one per flare, the one overlapping the peak HXR flux time).
The bin size (0.3) was taken larger than the average error (0.23).}
\label{fig:HistoDg}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=10cm]{HistoFr.ps}
\caption{Histogram of flux ratios.
Crosshatshed: ``Peak Time'' events. Solid black: ``Whole Peak'' events. Solid gray: ``Whole Peak'' events (only one per flare, the one overlapping the peak HXR flux time).
The bin size (0.5) is much larger than the typical error.}
\label{fig:HistoFr}
\end{figure}
\begin{table}
\centering
\caption{Consistency/inconsistency of $\Delta\gamma$ with zero: The numbers on the {\it left} of the ``$|$''
are the number of cases where $\Delta\gamma - n \cdot \sigma_{\Delta\gamma}>$0,
and numbers on the right are events where $\Delta\gamma + n \cdot \sigma_{\Delta\gamma}<$0.
The total number of each type of event is given in parenthesis at the top of each column.}
\begin{tabular}{c c c c}
\hline
Event type & ``Peak Times'' & ``Whole Peaks'' & main ``Whole Peak'' \\
& (33 events) & (87 events) & of each flare \\
& & & (37 events) \\
\hline
$n$=0 & 17$|$16 & 54$|$33 & 21$|$16 \\
$n$=1, ``1-sigma results'' & 10$|$15 & 34$|$19 & 12$|$11 \\
$n$=2, ``2-sigma results'' & 5$|$4 & 16$|$8 & 6$|$4 \\
$n$=3, ``3-sigma results'' & 2$|$1 & 5$|$2 & 1$|$0 \\
\hline
\end{tabular}
\label{table:DgPosNeg}
\end{table}
\begin{table}
\centering
\caption{Consistency/inconsistency of F$_\mathrm{r}$ with unity:
Events with F$_{\mathrm{r}} - n \cdot \sigma_{\mathrm{F}_{\mathrm{r}}}>$1 are left of the ``$|$'', and events with F$_{\mathrm{r}} + n \cdot \sigma_{\mathrm{F}_{\mathrm{r}}}<$1 are at the right. }
\begin{tabular}{c c c c}
\hline
Event type & ``Peak Times'' & ``Whole Peaks'' & main ``Whole Peak'' \\
& (33 events) & (87 events) & of each flare \\
& & & (37 events) \\
\hline
$n$=0 & 20$|$13 & 41$|$46 & 21$|$16 \\
$n$=1, ``1-sigma results'' & 20$|$13 & 40$|$39 & 21$|$15 \\
$n$=2, ``2-sigma results'' & 19$|$12 & 32$|$34 & 16$|$13 \\
$n$=3, ``3-sigma results'' & 19$|$11 & 24$|$32 & 12$|$12 \\
\hline
\end{tabular}
\label{table:Fr}
\end{table}
Roughly 25\% of ``Peak time'' events (8 out of 33) have spectral index differences consistent with zero ({\it i.e.}, $\Delta\gamma$ within 1-$\sigma$ of zero; Table \ref{table:DgPosNeg}).
This ratio increases to $\approx$40\% for ``Whole Peak'' events (34/87).
No ``Peak time'' event (0/33, Table \ref{table:Fr}) has a flux ratio consistent with unity, and only $\approx$10\% of the ``Whole peak'' events do (8/87; this fraction is almost reduced to zero (1/37) when considering the strongest ``Whole peak'' events of each flare ({\it i.e.} those encompassing the ``Peak time'')).
Table \ref{table:Fr} seem to suggest that leading footpoints might have more flux during ``Peak time'' events, but the result is not statistically significant, and will not be further discussed.
Figure \ref{fig:HistoDg} and Table \ref{table:DgPosNeg} show that there is no statistically significant preference for the leading footpoint to be either harder or softer than the trailing one, during either ``Peak Times'' or ``Whole Peaks''.
The deficit of F$_{50,\mathrm{r}}\approx$1 events during ``Peak Times'' is particularly clear in Figures \ref{fig:main} (bottom left) and \ref{fig:HistoFr} (only 1 out of 33 events is within 10\% of unity flux ratio, and only 5 out of 33 within 20\%).
This greater footpoint asymmetry during times of peak HXR fluxes than during whole HXR peaks could indicate that individual particle acceleration episode occur preferentially in one direction of the loop at any given time, but that, on the average, particles tend to be accelerated in both directions equally.
\begin{figure}
\centering
\includegraphics[width=10cm]{DgVsFr2.ps}
\caption{F$_{50,\mathrm{strong}}/F_{50,\mathrm{weak}}$ vs. $\gamma_{\mathrm{strong}}-\gamma_{\mathrm{weak}}$, where {\it strong} ({\it weak}) denotes the footpoint with the strongest (weakest) 50-keV flux, respectively.
The dashed line marks zero spectral index difference.
There are 23 out of 33 (79\%) ``Peak flux'' events and 53 out of 87 (61\%) ``Whole peak'' events which lie below the dashed line
(see also Table \ref{table:ColumnDensityModelCheck}.).
}
\label{fig:DgVsFr2}
\end{figure}
\begin{table}
\centering
\caption{Column density model: (agreement/disagreement).}
\begin{tabular}{c c c c}
\hline
Event type & ``Peak Times'' & ``Whole Peaks'' & All together \\
\hline
All events & 30.3\% (10/33)& 39.1\% (34/87)& 36.7\% (44/120)\\
1-sigma results & 24\% (6/25) & 42.9\% (21/49)& 36.5\% (27/74)\\
2-sigma results & 28.6\% (2/7) & 40.0\% (8/20) & 37.0\% (10/27)\\
3-sigma results & 33\% (1/3) & 60.0\% (3/5) & 50.0\% (4/8) \\
\hline
\end{tabular}
\label{table:ColumnDensityModelCheck}
\end{table}
We checked in a simple way the agreement of our data set with the theory presented in Appendix \ref{appendix:ColumnDensity}:
that spectral index differences $\Delta\gamma$ between flare footpoints might be due to differences in column depths in asymmetric loops.
This effect, assuming equal distributions of electrons are accelerated in both directions of the loop, results in the footpoint having the softer spectrum also having the most flux.
As shown in Figure \ref{fig:DgVsFr2} and Table \ref{table:ColumnDensityModelCheck}, the reverse happens most of the time,
{\it i.e.} better supporting a ``magnetic mirroring'' type of effect (see {\it e.g.} recent work by \opencite{Schmahl2007}, and references therein).
Furthermore, the best candidate in support of the column density difference model is the 23 July 2002 flare (which agrees at the ``3-$\sigma$ level'') actually hits a snag when one
considers the amount of 50-keV flux (predicted by theory) out of the footpoints ({\it i.e.} emitted somewhere along the legs of the loop, before reaching the imaged footpoints): as explained in Appendix \ref{appendix:ColumnDensity},
this 50-keV emission should have been observable.
Moreover, the presence of large flux ratios F$_{\mathrm{r}}$ (2 or above, $\frac{1}{2}$ or below) also lead us to believe that this theory cannot be a dominant factor, at least for reasonable values of leg column densities (see Appendix \ref{appendix:ColumnDensity}).
There are several altenatives to explain footpoint asymmetries:
(a) {\it Asymmetrical acceleration:} The strong footpoint asymmetries, particularly during HXR peak times, suggest that it is the acceleration mechanism itself which could be asymmetrical.
If the acceleration process actually took place in the chromospheric footpoints (as opposed to high in the corona), one would expect asymmetries, as both acceleration processes could in principle be independent from one another.
(b) {\it Non-uniform target ionization:} It is conceivable that both chromospheric footpoints have different ionization structure,
{\it i.e.} considering the simplified step model of \inlinecite{Brown1973} or \inlinecite{Kontar2002}, that the column density required before reaching the lower-chromospheric regions of neutral atoms
is different in both legs of the loop, perhaps due to some prior heating of only one of the footpoints.
Modeling and comparison with observations are required to further this idea.
(c) The best candidate mechanism to explain footpoint emission asymmetries is magnetic mirroring, as discussed in \inlinecite{Aschwanden1999}'s
{\it trap+precipitation} model \cite{Melrose1976}: the magnetic field converges more rapidly in one of the footpoints, and particles are mirrored before they reach the dense lower regions (see also \opencite{Schmahl2007}).
Our data corroborate better that scenario than the column density asymmetry model.
The effects of {\it photospheric albedo} \cite{Bai1978} or {\it return currents} \cite{Zharkova2006} might reinforce any asymmetry observed in footpoint photon spectra, but only if the accelerated electron distributions started out different.
\subsection{Correlation Table:}\label{sect:obs:corr}
Correlation coefficients have been computed for our data, and are displayed in Table \ref{table:corr}.
\begin{table}
\rotatebox{90}{
\begin{minipage}[c]{15cm}
\begin{tiny}
\centering
\begin{tabular}{c c c c c c c c}
\hline
& Duration & 50-keV & 50-keV & & & \\
& [s] & flux ratio & flux & $\Delta\gamma$ & $\vert\Delta\gamma\vert$ & $\bar{\gamma}$ \\
&&$\ln \left( \frac{F_{50,1}}{F_{50,2}} \right)$&$\ln (F_{50,1}+F_{50,2})$&&&&\\
\hline
Duration [s] & -/1.00 & -/-0.08 & -/-0.01 & -/-0.07 & -/-0.11 & -/-0.21 \\
$\ln \left( \frac{F_{50,1}}{F_{50,2}} \right)$ & & 1.00/1.00 & 0.06/0.12 & {\bf -0.53}/-0.23 & 0.45/0.09 & 0.32/0.22 \\
$\ln (F_{50,1}+F_{50,2})$ & & & 1.00/1.00 & 0.24/0.11 & -0.22/-0.13 & -0.23/-0.24 \\
$\Delta\gamma$ & & & & 1.00/1.00 & -0.42/0.16 & -0.41/-0.08 \\
$\vert\Delta\gamma\vert$ & & & & & 1.00/1.00 & {\bf 0.50}/0.49 \\
$\bar{\gamma}$ & & & & & & 1.00/1.00 \\
Longitude & & & & & & \\
$\vert$Longitude$\vert$ & & & & & & \\
Latitude & & & & & & \\
$\vert$Latitude$\vert$ & & & & & & \\
FP separation & & & & & & \\
FP angle & & & & & & \\
$\ln (\mathrm{GOES flux})$ & & & & & & \\
\hline
& & & & & FP & FP & GOES \\
& Longitude & $\vert$Longitude$\vert$ & Latitude & $\vert$Latitude$\vert$ & separation & angle & class \\
\hline
Duration [s] & -/0.02 & -/-0.08 & -/-0.03 & -/0.10 & -/{\bf 0.52} & -/-0.05 & -/0.22 \\
$\ln \left( \frac{F_{50,1}}{F_{50,2}} \right)$ & 0.15/0.05 & -0.24/-0.12 & -0.01/0.07 & 0.43/0.10 & -0.12/-0.07 & -0.07/0.01 & 0.16/0.12 \\
$\ln (F_{50,1}+F_{50,2})$ & 0.26/-0.06 & 0.12/0.17 & 0.20/0.01 & -0.25/-0.14 & 0.04/0.00 & 0.22/0.20 & {\bf 0.88}/{\bf 0.71} \\
$\Delta\gamma$ & -0.01/-0.08 & 0.33/0.31 & -0.20/-0.19 & -0.12/-0.18 & -0.18/-0.03 & 0.05/0.25 & 0.11/0.03 \\
$\vert\Delta\gamma\vert$ & -0.01/-0.02 & -0.25/0.12 & 0.12/-0.20 & 0.19/0.13 & 0.01/-0.18 & -0.04/0.08 & -0.09/-0.07 \\
$\bar{\gamma}$ & 0.29/0.31 & -0.24/-0.16 & 0.06/-0.03 & -0.06/0.11 & -0.01/-0.42 & -0.26/-0.02 & -0.12/-0.22 \\
Longitude & 1.00 & -0.04 & 0.17 & -0.06 & -0.02 & 0.09 & 0.12 \\
$\vert$Longitude$\vert$ & & 1.00 & -0.27 & -0.08 & 0.13 & 0.13 & 0.07 \\
Latitude & & & 1.00 & -0.27 & -0.08 & 0.25 & -0.03 \\
$\vert$Latitude$\vert$ & & & & 1.00 & -0.18 & -0.12 & 0.02 \\
FP separation & & & & & 1.00 & -0.15 & 0.22 \\
FP angle & & & & & & 1.00 & 0.13 \\
$\ln$(GOES flux) & & & & & & & 1.00 \\
\hline
\end{tabular}
\end{tiny}
\end{minipage}
}
\caption{Table of correlation coefficient between footpoint parameters: Where there are two numbers separated by a slash, the number to the left applies to ``Peak Times'',
and the one to the right applies to ``Whole Peaks''. Correlation coefficients with magnitude greater than 0.5 are in bold. }
\label{table:corr}
\end{table}
We found the following:
\begin{itemize}
\item There is some degree of anti-correlation (-0.53, with the 95\% confidence interval being [-0.74,-0.22]) between $\Delta\gamma$ and F$_{\mathrm{r}}$, for ``Peak Times'' events.
This is a consequence of there having more events in the {\it upper left} and {\it lower right} quadrants of the lower left panel of Figure \ref{fig:main}, and has already been discussed in Section \ref{sect:obs:spectral}.
\item $\bar{\gamma}$ and $\vert\Delta\gamma\vert$ seem also slightly correlated: indicating that $\vert\Delta\gamma\vert$ is larger when the flare is softer.
Upon closer examination, it appears softer spectra to be a simple case of softer spectra having larger errors.
\item An unexpected correlation --albeit weak (the 95\% confidence interval is [0.32, 0.67])-- was found between footpoint separation and event duration, for ``Whole Peaks'' events.
\begin{figure}
\centering
\includegraphics[width=10cm]{durationseparation.ps}
\caption{``Whole Peak'' duration vs. footpoint separation. The error bars are small and were not omitted.
The gray line is a linear fit (using the bisector method) to the data.
It yields a power-law slope of 0.6$\pm$0.1.
}
\label{fig:durationseparation}
\end{figure}
From Figure \ref{fig:durationseparation}, it seems that we have
\begin{equation}
\mathrm{(HXR \,\,\, burst \,\,\, duration)} \approx \mathrm{(footpoint \,\,\, separation)^2},
\end{equation}
or, assuming semi-circular loops:
\begin{equation}
\mathrm{(HXR \,\,\, burst \,\,\, duration)} \approx \mathrm{(loop \,\,\, length)^2}
\end{equation}
It seems that the longer the loops, the longer the HXR peak will last.
The interpretation is not yet clear.
It could be a simple case of larger loops needing more time to evolve than the short ones during the flaring process,
or another case of the ``big flare syndrome'': everything is bigger in larger flares.
\item Excellent correlation between ``Peak Times'' total flux and GOES class (95\% confidence interval for correlation coefficient: 0.77--0.94), less good for ``Whole Peaks'' total flux and GOES class (95\% confidence interval for correlation coefficient: 0.57--0.81).
\end{itemize}
\begin{figure}
\centering
\includegraphics[width=10cm]{GOESvsF50.ps}
\caption{GOES maximum flux in the 1--8{\AA} band vs. peak 50-keV flux of both footpoints, with 1-$\sigma$ errors.
The solid line is a power-law linear fit to the data.
The dotted line is another power-law fit to the data, using the bisector method.}
\label{fig:GOESvsF50tot}
\end{figure}
Figure \ref{fig:GOESvsF50tot} shows the clear correlation between event's maximum GOES 1--8{\AA} flux and the
total HXR flux F$_{50}$, for ``Peak Time'' events. Fitting a power-law yields:
\begin{equation}
\mathrm{F}_{50} = A \cdot \mathrm{F}_{\mathrm{GOES},1-8\mathrm{\AA}}^\alpha
\end{equation}
where $A=(4.7\pm0.3) \times 10^3$ and $\alpha=0.8\pm0.1$ when F$_{50}$ is in photons s$^{-1}$ cm$^{-2}$ keV$^{-1}$ and F$_{\mathrm{GOES},1-8\mathrm{\AA}}$ in W m$^{-2}$.
The {\it bisector} method \cite{Isobe1990} is more relevant when the variables are independent,
in which case we have $A=10^{4.3\pm0.2}$ and $\alpha=0.97\pm0.05$.
(Of course, ``Whole HXR Peaks'' events typically lie below this solid line, with a wide scatter.)
This good correlation can be interpreted as a direct consequence of the Neupert effect \cite{Neupert1968,Dennis1985}: the larger the amount of non-thermal energy (approximated by F$_{50}$), the larger the amount of
thermal energy (approximated by the GOES SXR) (since F$_{50}$ is a power-law normalization factor, it describes equally well the amount of non-thermal electrons of lower energies, which contain most of the non-thermal power).
Similar correlations have been reported before (see {\it e.g.} \opencite{Marina2005} and references therein.)
\section{Summary and Conclusion}\label{sect:ccl}
The following is a compilation of our results, and can be used as a list of contraints for any flare and particle acceleration theories:
(1) The total footpoint 50-keV flux correlates remarkably well with the GOES maximum 1--8{\AA} flux. The relationship is fairly linear.
(2) There is no statistically significant difference in our sample between ``leading'' and ``trailing'' footpoints, as regards asymmetries.
(3) Flares are mostly located at $\pm$15 degrees of solar latitude, and flare parameters have no marked longitudinal dependence.
(4) Spherical separation between footpoints seem not to correlate with any of the other parameters examined, with the surprising exception of HXR burst duration, where a weak correlation was found.
This seems to indicate that longer loops produce longer HXR peaks, probably because the magnetic disturbance and particle acceleration last longer in long loops than in short ones.
(5) Flare footpoint spectral indices $\gamma$ are seldom below $\approx$2.4 (1 case out of 172).
``Peak times'' are generally harder than ``Whole Peak'' intervals, a natural consequence of the commonly observed soft-hard-soft behaviour of flares.
(6) $\approx$25\% (``Peak times'') to $\approx$40\% (``Whole Peaks'') of double footpoint flares have spectral index differences $\Delta\gamma$ consistent with zero.
$\Delta\gamma$ can reach 0.6, and only rarely goes beyond.
The amplitude of $\Delta\gamma$ is uncorrelated with flare GOES class.
(7) 50-keV footpoint flux ratios are never quite unity, are typically between 1 and 2, and only seldom go beyond 3. This result could be due to observational bias.
(8) The asymmetric loop model, where a column density difference is responsible for the difference in spectral index and flux between HXR footpoints, cannot explain a majority of our observations.
It is therefore not a dominant factor.
(9) The greatest asymmetry being around ``Peak Times'' further suggests that magnetic reconfiguration is greatest at those times.
|
\section{Introduction}
The question about how a piece of information (a virus, a rumor, an opinion, etc.,) is globally spread over a network, and which ingredients are necessary to achieve such a success, has motivated much research recently. The reason behind this interest is that identifying key aspects of spreading phenomena facilitates the prevention (e.g., minimizing the impact of a disease) or the optimization (e.g. the enhancement of viral marketing) of diffusion processes that can reach system wide scales. In the context of political protest or social movements, information diffusion plays a key role to coordinate action and to keep adherents informed and motivated \cite{gonzalez2011}. Understanding the dynamics of such diffusion is important to locate who has the capability to transform the emission of a single message into a global information cascade, affecting the whole system. These are the so-called ``privileged or influential spreaders''. Beyond purely sociological aspects, some valuable lessons might be extracted from the study of this problem. For instance, current viral marketing techniques (which capitalizes on online social networks) could be improved by encouraging customers to share product information with their acquaintances. Since people tend to pay more attention to friends than to advertisers, targeting privileged spreaders at the right time may enhance the efficiency of a given campaign.
The prominence (importance, popularity, authority) of a node has however many facets. From a static point of view, an authority may be characterized by the number of connections it holds, or the place it occupies in a network. This is the idea put forward in \cite{gupte2011finding}, where the authors seek the design of efficient algorithms to detect particular (sub)graph structures: hierarchies and tree-like structures. Turning to dynamics, a node may become popular because of the attention it receives in short intervals of time \cite{ratkiewicz2010characterizing} --but that is a rather volatile way of being important, because it depends on activity patterns that change in the scale of hours or even minutes. A more lasting concept of influence comprises both a topological --enduring-- ingredient and the dynamics it supports; this is the case of Centola's ``reinforcing signals'' \cite{centola10science} or the $k$-core \cite{kitsak2010identification}, which we follow here.
In this paper, we approach the problem of influential spreaders taking into consideration data from the Spanish ``15M movement'' \cite{borge2011structural}. This pacific civil movement is an example of the social mobilizations --from the ``Arab spring'' to the ``Occupy Wall-Street'' movement -- that have characterized 2011. Although whether OSNs have been fundamental instruments for the successful organization and evolution of political movements is not firmly established, it is increasingly evident \cite{borge2011structural} that at least they have been nurtured mainly in OSNs (Facebook, Twitter, etc.) before reaching classic mass media. Data from these grassroots movements --but also from less conflictive phenomena in the Web 2.0-- provide a unique opportunity to observe system-wide information cascades. In particular, paying attention to the network structure allows for the characterization of which users have outstanding roles for the success of cascades of information. Our results complement some previous findings regarding dynamical influence both at the theoretical \cite{kitsak2010identification} and the empirical \cite{gonzalez2011} levels. Besides, our analysis of activity cascades reveals distinctive traits in different phases of the protests, which provides important hints for future modeling efforts.
\section{Data: a networked view of the 15M movement}
\label{sec:data}
The ``15M movement'' is a still ongoing civic initiative with no party or union affiliation that emerged as a reaction to perceived political alienation and to demand better channels for democratic representation. The first mass demonstration, held on Sunday May 15 ($D$ from now on), was conceived as a protest against the management of the economy in the aftermath of the financial crisis. After the demonstrations on day $D$, hundreds of participants decided to continue the protests camping in the main squares of several cities (Puerta del Sol in Madrid, Pla\c ca de Catalunya in Barcelona) until May 22, the following Sunday and the date for regional and local elections.
From a dynamical point of view, the data used in this study are a set of messages (tweets) that were publicly exchanged through {\em www. twitter.com}. The whole time-stamped data collected comprises a period of one month (between April 25th, 2011 at 00:03:26 and May 26th, 2011 at 23:59:55) and it was archived by {\em Cierzo Development Ltd.}, a start-up company. To filter out the whole sample and choose only those messages related to the protests, 70 keywords ({\em hashtags}) were selected, those which were systematically used by the adherents to the demonstrations and camps. The final sample consists of 535,192 tweets. On its turn, these tweets were generated by 85,851 unique users (out of a total of 87,569 users of which 1,718 do not show outgoing activity, i.e., they are only receivers). See \cite{web15m} for more details.
Twitter is most frequently used as a broadcasting platform. Users subscribe to what other users say building a ``who-listens-to-whom'' network, i.e., that made up of followers and followings in Twitter. This means that any emitted message from a node will be immediately available to anyone following him, which is of utmost importance to understand the concept of activity cascade in the next sections. Such relationships offer an almost-static view of the relationships between users, the ``follower network'' for short. To build it, data for all the involved users were scrapped directly from {\em www.twitter.com}. The scrap was successful for the 87,569 identified users, for whom we also obtained their official list of followers restricted to those who had some participation in the protests. The resulting structure is a directed network, direction indicates who follows who in the online social platform. In practice, we take this underlying structure as completely static (does not change through time) because its time scale is much slower, i.e., changes occur probably in the scale of weeks and months. In-degree $k_{in}$ expresses the amount of users a node is following; whereas out-degree represents the amount of users who follow a node. This network exhibits a high level of reciprocity: a typical user holds many reciprocal relationships (with other users who the node probably knows personally), plus a few unreciprocated nodes which typically point at hubs.
\begin{figure}[!th]
\begin{center}
\includegraphics[width=\columnwidth,clip=0]{Figure1.eps}
\caption{(Color online) Temporal evolution of the activity in the online social network. In green, the proportion of nodes that had shown some activity at a certain time $t$. In yellow, the cumulative proportion of emitted messages as a function of time. Note that the two lines evolve in almost the same way. According to this evolution, we have distinguished two sub-periods: one of them characterized as ``slow growth'' due to the low activity level and the other one tagged as ``explosive'' or "bursty" due to the intense information traffic within it.}
\label{growth}
\end{center}
\end{figure}
The main topological features of the follower network fit well in the concept of ``small-world'' \cite{watts98}, i.e., low average shortest path length and high clustering coefficient. Furthermore, both in- and out-degree distribute as a power-law, indicating that connectivity is extremely heterogenous. Thus, the network supporting users' interactions is scale-free with some rare nodes that act as hubs \cite{barabasi99}.
\begin{figure*}[!th]
\centering
\includegraphics[width=0.85\linewidth,clip=0]{Figure2.eps}
\caption{(Color online) The figure illustrates the concept of cascade that is used throughout this article. User 1 emits a message at time $t$, and all of his followers automatically receive it. Thus, they are already counted as part of the cascade (small red circles). One of his followers (user 2, big blue node), driven by the previous message, decides himself to participate at time $t+\Delta t$, posting a message himself. A second set of followers are included in the cascade. Finally, a third node (user 3, big green circle) joins in and spreads the cascade further at time $t+2\Delta t$. A node can not be counted twice, note for example that user 4 is also following node 3. Many nodes remain unaffected, because they are not connected to any of the spreaders. The final size of the cascade is $\frac{N_{c}}{N} = \frac{22}{34}$; the success of the cascade largely depends on the capacity to contact a ``leader'' or ``privileged spreader'', i.e., a hub to whom many people listens and who decides to participate. The interesting point, however, is that the number of spreaders needed to attain such success is very low (3), and over 50\% of the cascade is triggered by just one of them.}
\label{example}
\end{figure*}
\section{Methods}
\label{sec:methods}
\subsection{Activity cascades}
An activity cascade --or simply ``cascade'', for short--, starting at a \emph{seed}, occurs whenever a piece of information --or replies to it-- are (more or less unchanged) repeatedly forwarded towards other users. If one of those who ``hear'' the piece of information decides to reply to it, he becomes a \emph{spreader}, otherwise he remains as a mere \emph{listener}. The cascade becomes global if the final number of affected users $N_{c}$ (including the set of spreaders and listeners, plus the seed) is comparable to the size of the whole system $N$. Intuitively, the success of an activity cascade greatly depends on whether spreaders have a large set of followers or not (Figure \ref{example}); remarkably, the seed is not necessarily very well connected. This fact highlights the entanglement between dynamics and the underlying (static) structure.
Note that the previous definition is too general to attain an \emph{operative} notion of cascade. One possibility is to leave time aside, and consider only identical pieces of information traveling across the topology (a {\em retweet}, in the Twitter jargon). This may lead to inconsistencies, such as the fact that a node decides to forward a piece of information long after receiving it (perhaps days or weeks). It is impossible to know whether his action is motivated by the original sender, or by some exogenous reason, i.e., invisible to us. One may, alternatively, take into consideration time, thus considering that, regardless of the exact content of a message, two nodes belong to the same cascade as consecutive spreaders if they are connected (the latter follows the former) and they show activity within a certain (short) time interval, $\Delta t$. The probability that exogenous factors are leading activation is in this way minimized. Also, this concept of cascade is more inclusive, regarding dialogue-like messages (which, we emphasize, are typically produced in short time spans). This scheme exploits the concept of spike train from neuroscience, i.e., a series of discrete action potentials from a neuron taken as a time series. At a larger scale, two brain regions are identified as functionally related if their activation happens in the same time window. Consequently, message chains are reconstructed assuming that activity is contagious if it takes place in short time windows.
\begin{figure*}[!th]
\centering
\includegraphics[width=0.9\linewidth,clip=0]{Figure3.eps}
\caption{(Color online) Upper panels (\emph{a,b,c}): Cascade size probability distributions for the different periods considered. Lower panels (\emph{d,e,f}): Probability distributions of spreaders involved in the cascades for the same periods. The exact periods considered in the analyses are indicated at the top of each panel. See the text for further details.}
\label{fig3}
\end{figure*}
We apply the latter definition to explore the occurrence of information cascades in the data. In practice, we take a seed message posted by $i$ at time $t_{0}$ and mark all of $i$'s followers as listeners. We then check whether any of these listeners showed some activity at time $t_{0}+\Delta t$. This is done recursively until no other follower shows activity, see Figure \ref{example}. In our scheme, a node can only belong to one cascade; this constraint introduces a bias in the measurements, namely, two nodes sharing a follower may show activity at the same time, so their follower may be counted in one or another cascade (with possible important consequences regarding average cascades' size and penetration in time). To minimize this degeneration, we perform calculations for many possible cascade configurations, randomizing the way we process data. We distinguish information cascades (or just cascades, for short) from spreader-cascades. In information cascades we count any affected user (listeners and spreaders), whereas in spreader-cascades only spreaders are taken into account.
We measure cascades and spreader-cascades size distributions for three different scenarios: one in which the information intensity is low (slow growth phase, from $D-20$ to $D-10$), one in which activity is bursty (explosive phase, $D-2$ to $D+6$) and one that considers all available data (which spans a whole month, and includes the two previous scenarios plus the time in-between, $D-20$ to $D+10$). Figure \ref{growth} illustrates these different periods. The green line represents the cumulative proportion of nodes in the network that had shown some activity, i.e., had sent at least one message, measured by the hour. We tag the first 10 days of study as ``slow growth'' because, for that period, the amount of active people grew less than 5\% of the total of users, indicating that recruitment for the protests was slow at that time. The opposite arguments apply in the case of the bursty or ``explosive'' phase: in only 8 days the amount of active users grew from less than 10\% up to over an 80\%. The same can be said about global activity (in terms of the total number of emitted directed messages --the activity network), which shows an almost exact growth pattern. Besides, within the different time periods --slow growth, explosive and total--, different time windows have been set to assess the robustness of our results. Our proposed scheme relies on the contagious effect of activity, thus large time windows, i.e., $\Delta t > 24$ hours, are not considered.
\subsection{$k$-shell decomposition}
The $k$-core decomposition of a network consists of identifying particular subsets of the network, called $k$-cores, each obtained by recursively removing all the vertices of degree less than $k$, where $k = k_{in} + k_{out}$ indicates the total number of in- and out-going links of a node, until all vertices in the remaining graph have degree at least $k$. In the end, each node is assigned a natural number (its coreness), the higher the coreness the closer a node is to the nucleus or core of the network. The main advantage of this centrality measure is, in front of other quantities, its low computational cost that scales as $O(N+E)$, where $N$ is the number of vertices of the graph and $E$ is the number of links it contains \cite{alvarez2008k}. This decomposition has been successfully applied in the analysis of the Internet and the Autonomous Systems structure \cite{alvarez2008k,carmi2007model}. In the following section, we will use the $k$-core decomposition as a means to identify influence in social media. In particular, we discuss which, degree or coreness, is a better predictor of the extent of an information cascade.
\begin{figure*}[!th]
\centering
\includegraphics[width=0.85\linewidth,clip=0]{Figure4.eps}
\caption{(Color online) Left upper panel: average spreading capacity (with respect to the system size) of nodes grouped according to their $k$-core. $\frac{N_{c}}{N}$ grows with coreness, but the explosive period (red squares) evidences a much less clear tendency, with many fluctuations and a lower overall spreading capacity if compared to the slow growth period (black circles). Left lower panel: The same information is showed as a function of the degree. Again, the slow growth period is the best one at predicting the extent of a cascade. Interestingly, average cascades for highest degrees outperform those triggered by highest $k$-core nodes by an order of magnitude. See main text for discussion on this aspect. Right panels show the $k$-core and degree distributions, i.e., how many nodes belong to each class. Note that the highest core contains over 1000 users.}
\label{coredegree}
\end{figure*}
\section{Results}
\label{sec:results}
The upper panels ($a,b,c$) of Figure \ref{fig3} reflect that a cascade of a size $O(N)$ can be reached at any activity level (slow growth, explosive or both). As expected, these large cascades occur rarely as the power-law probability distributions evidence. This result is robust to different temporal windows up to 24h.
In contrast, lower ($d,e,f$) panels show significant differences between periods. Specifically, the distribution of involved spreaders in the different scenarios changes radically from the ``slow growth'' phase (Figure \ref{fig3}d) to the ``explosive'' period (Figure \ref{fig3}f); the distribution that considers the whole period of study just reflects that the bursty period (in which most of the activity takes place) dominates the statistics. The importance of this difference is that one may conclude that, to attain similar results a proportionally much smaller amount of spreaders is needed in the slow growth period. Going to the detail, however, it seems clear (and coherent with the temporal evolution of the protests, Fig. \ref{growth}) that although cascades in the slow period (panel a) affect as much as $N/2$ of the population, the system is in a different dynamical regime than in the explosive one: indeed, distributions suggest that there has been a shift from a subcritical to a supercritical phase.
The previous conclusions raise further questions: is there a way to identify ``privileged spreaders''? Are they placed randomly throughout the network's topology? Or do they occupy key spots in the structure? And, will these influential users be more easily detected in a bursty period (where large cascades occur more often)? In what context will influential spreaders single out? To answer these questions, we capitalize on previous work suggesting that centrality (measured as the $k$-core) enhances the capacity of a node to be key in disease spreading processes \cite{kitsak2010identification}. The authors in \cite{kitsak2010identification} discussed whether the degree of a node (its total number of neighbors, $k$) or its $k$-core (a centrality measure) can better predict the spreading capabilities of such node. Note that the $k$-shell decomposition splits a network in a few levels (over a hundred), while node degrees can range from one or two up to several thousands.
We have explored the same idea, but in relation to activity cascades which are the object of interest here. The upper left panel of Fig.\ \ref{coredegree} shows the spreading capabilities as a function of classes of $k$-cores. Specifically, we take the seed of each particular cascade and save its coreness and the final size of the cascade it triggers. Having done so for each cascade, we can average the success of cascades for a given core number. Remarkably, for every scenario under consideration (slow, explosive, whole), a higher core number yields larger cascades. This result supports the ideas developed in \cite{kitsak2010identification}, but it is at odds with those reported in \cite{borge2012absence}, which shows that the $k$-core of a node is not relevant in rumor dynamics.
Exactly the same conclusion (and even more pronounced) can be drawn when considering degree (lower left panel), which appears to be in contradiction with the mentioned previous evidence \cite{kitsak2010identification}.
At a first sight, our findings seem to point out that if privileged spreaders are to be found, one should simply identify the individuals who are highly connected. However, this procedure might not be the best choice. The right panels in Figure \ref{coredegree} show the $k$-core (upper) and degree (lower) distributions, indicating the number of nodes which are seeds at one time or another, classified in terms of their coreness or degree. Unsurprisingly, many nodes belong to low cores and have low degrees. The interest of these histograms lies however in the tails of the distributions, where one can see that, while there are a few hundred nodes in the high cores (and even over a thousand in the last core), highest degrees account only for a few dozen of nodes. In practice, this means that by looking at the degree of the nodes, we will be able to identify quite a few influential spreaders (the ones that produce the largest cascades). However, the number of such influential individuals are far more than a few. As a matter of fact, high cascading capabilities are distributed over a wider range of cores, which in turn contain a significant number of nodes. Focusing on Fig.\ \ref{coredegree}, note that triggering cascades affecting over $10^{-2}$ of the network's population demands nodes with $k \ge 10^{3}$. Checking the distribution of degrees (right-hand side), it is easy to see that an insignificant amount of nodes display such degree range. In the same line, we may wonder what it takes to trigger cascades affecting over $10^{-2}$ of the network's population, from the $k$-core point of view. In this case, nodes with $k$-core around 125 show such capability. A quick look at the core distribution yields that over 1500 nodes accomplish these conditions, i.e., they belong to the 125th $k$-shell or higher.
We may now distinguish between scenarios in Figure \ref{coredegree}. While any of the analyzed periods shows a growing tendency, i.e., cascades are larger the larger is the considered descriptor, we highlight that it is in the slow growth period (black circles) where the tendency is more clear, i.e., results are less noisy. Between the other two periods, the explosive one (red squares) is distinctly the less robust, in the sense that cascade sizes oscillate very much across $k$-cores, and the final plot shows a smaller slope than the other two. This subtle fact is again of great importance: it means that during ``information storms'' a large cascade can be triggered from anywhere in the network (and, conversely, small cascades may have begun in important nodes). The reason for this is that in periods where bursty activity dominates the system suffers ``information overflow'', the amount of noise flattens the differences between nodes. For instance, in these periods a node from the periphery (low coreness) may balance his unprivileged situation by emitting messages very frequently. This behavior yields a situation in which, from a dynamical point of view, nodes become increasingly indistinguishable. The plot corresponding to the whole period analyzed (green triangles) lies consistently between the other two scenarios, but closer to the relaxed period. This is perfectly coherent, the study spans for 30 days and the explosive period represents only 25\% of it, whereas the relaxed period stands for over 33\%. Furthermore, those days between $D-10$ and $D-2$, and beyond $D+6$, resemble the relaxed period as far as the flow of information is concerned.
\section{Conclusions}
\label{sec:conclusions}
Online social networks are called to play an ever increasing role in shaping many of our habits (be them commercial or cultural) as well as in our position in front of political, economical or social issues not only at a local, country-wide level, but also at the global scale. It is thus of utmost importance to uncover as many aspects as possible about topological and dynamical features of these networks. One particular aspect is whether or not one can identify, in a network of individuals with common interests, those that are influentials to the rest. Our results show that the degree of the nodes seems to be the best topological descriptor to locate such influential individuals. However, there is an important caveat: the number of such privileged seeds is very low as there are quite a few of these highly connected subjects. On the contrary, by ranking the nodes according to their $k$-core index, which can be done at a low computational cost, one can safely locate the (more abundant in number) individuals that are likely to generate large (near to) system-wide cascades. The results here presented also lead to a surprising conclusion: periods characterized by explosive activity are not convenient for the spreading of information throughout the system using influential individuals as seeds. This is because in such periods, the high level of activity --mainly coming from users which are badly located in the network-- introduces noise in the system. Consequently, influential individuals lose their unique status as generators of system wide cascades and therefore their messages are diluted.
On more general grounds, our analysis of real data remarks the importance of empirical results to validate theoretical contributions. In particular, Fig. \ref{coredegree}, together with the observations in \cite{borge2012absence}, raises some doubts about rumor dynamics as a good proxy to real information diffusion. We hypothesize that such models approach information diffusion phenomena in a too simplistic way, thus failing to comprise relevant mechanisms such as complex activity patterns \cite{barabasi2005origin,fernandez11,vazquez2007impact}.
Finally, although the underlying topology may be regarded as constant, any modeling effort should also contemplate the time evolution of the dynamics. Indeed, Fig. \ref{fig3} suggests that the system is in a sub-critical phase when activity level is low, and critical or supercritical during the explosive period. This is related to the rate at which users are increasingly being recruited as active agents, i.e. the speed at which listeners become spreaders.
\section*{Acknowledgments}
This work has been partially supported by MICINN through Grants FIS2008-01240 and FIS2011-25167, and by Comunidad de Arag\'on (Spain) through a grant to the group FENOL.
|
\section{Model and Methods}
\subsection{Formulation of Limited Path Percolation induced by disorder}
To be concrete, LPP in Ref.~\cite{Lopez-LPP}, where link failure corresponds to link removal,
is formulated in the following way: if a pair of
nodes $i$ and $j$ is connected through a shortest path of length $\ell_{ij}$ (number of links)
at $p=1$ (no links removed), and of length $\ell^{'}_{ij}(p)$ at $p<1$ (fraction $1-p$ of links removed) ,
$i$ and $j$ are considered reachable if
$\ell^{'}_{ij}(p)\leq \tau\ell_{ij}$, where $\tau$ is the tolerance
factor which lies on the range between 1 and $\infty$~\cite{Lopez-LPP}.
The number of reachable pairs, $S$, is measured through
\begin{equation}
S=\sum_{i\neq o} \theta(\tau\ell_{oi}-\ell^{'}_{oi})
\label{S}
\end{equation}
where $o$ is an ``origin'' node chosen to minimize its impact on $S$ (see below), and
$\theta$ is the Heaveside step function $\theta(x)=1$ is $x\geq 0$, and 0 otherwise.
$S$ depends on both $p$ and $\tau$, and the
LPP phase transition occurs for the combination of these parameters at the threshold of the relation $S\sim N$.
For fixed $\tau$, there is a threshold $p=\tilde{p}_c(\tau)$
at which $S\sim N$ is achieved. Alternatively, given $p\geq p_c$~\cite{fnpgtrpc}, there is a
$\tau=\tau_c(p)$ such that $S\sim N$. When $\tau\rightarrow\infty$, $\tilde{p}_c(\tau\rightarrow\infty)=p_c$.
For disorder induced LPP, the set up is similar, but instead of considering
each link to be kept with probability $p$, we
change each link weight from 1 to $w$ drawn from a random uncorrelated distribution $P_a(w)$
where $a$ is the disorder parameter (defined below)~\cite{reg-LPP}.
The path lengths change from $\ell_{ij}$ to $\ell^{'}_{ij}(a)$ where now the later
corresponds to the length of the optimal path between $i$ and $j$.
Reachability is then defined as in Ref.~\cite{Lopez-LPP}: $i$ and $j$ are considered
reachable if $\ell^{'}_{ij}(a)\leq \tau\ell_{ij}$. The number of node pairs $S$ that remain
reachable is a function of $a$, $\tau$, and $N$. We search for the LPP
threshold by imposing $S\sim N$. If a threshold exists
for a fixed tolerance $\tau$, there is a critical disorder $a=a_c(\tau)$ or,
vice versa, a critical tolerance $\tau=\tau_c(a)$ that depends on the disorder parameter $a$.
Another way to phrase this is to say that $\tau$ and $a$ are control parameters in disorder
induced LPP, as $\tau$ and $p$ are in regular LPP.
To consider a tolerance to the path length increase without considering an associated tolerance to the
path weight increase may at first seem arbitrary, but in fact
is well justified in that the overall weight of a path is asymptotically proportional to its length
under the conditions of disorder we study here. Hence, choosing path length tolerance does
not affect the qualitative nature of our results, and one choice of tolerance can be mapped
onto the other. In concrete terms, if we imagined disorder corresponding to something like
time of travel through a link, then total average travel time scales linearly with the trip distance.
The general algorithm used to measure disorder induced LPP is the following.
First, we select the ensemble of networks $\mathcal{G}$ of interest. In this article
we focus on 2-dimensional square lattices and Erd\H{o}s-R\'{e}nyi (ER) graphs~\cite{ER}.
Square lattices have no randomness, of course, but can be formally viewed as an ensemble
with one single realization.
For each network realization $G\in\mathcal{G}$, in which all links have weight 1,
we determine the path length $\ell_{oi}$ between nodes $o$ and $i$ for all $i\neq o$.
On the same network realization $G$ (the same nodes and links),
disorder is introduced by changing the weight of each link from 1 to $w_{ij}$. Subsequently,
the optimal paths between $o$ and all other nodes $i$ of $G$ are determined, and their lengths
$\ell^{'}_{oi}$ recorded. $S$ is calculated by using Eq.~(\ref{S}).
To determine path lengths, we use the Dijkstra algorithm~\cite{Cormen}.
\subsection{Disorder}
We consider disorder distributions $P_a(w)$ characterized by a single
disorder parameter which, for convenience, we label as $a$. Generally, we would expect
that disorder induced LPP changes as a function of the form of $P_a(w)$.
However, recent work~\cite{Chen} shows
that large classes of disorder distributions are essentially equivalent in the optimal path
problem, provided a certain characteristic length scale $\xi(a)$ associated with $P_a(w)$ is conserved
(see below).
In practical terms, this means that disorder distributions of different functional forms but with
equal $\xi(a)$ lead to optimal path distributions that scale in the same way~\cite{Chen}.
This result allows us to choose
a distribution that is convenient and well understood. Thus, we use
\begin{equation}
P_a(w)=(aw)^{-1},\qquad [w\in [1,e^a]]
\label{Pw}
\end{equation}
for which a large amount of research has been conducted in the context of the optimal path
problem~\cite{Cieplak,Porto,Braunstein,Wu,Ambegaokar}.
Determining $\xi$ for a given distribution is addressed in~\cite{Chen}, and we
return to this in the discussion of results.
\section{Results and Discussion}
To characterize LPP effectively, we first measure $\Theta(S|\tau,a,N)$, the distribution of sizes of
the cluster containing $o$, for networks of size $N$ with tolerance $\tau$
over disorder and network realizations.
Previously~\cite{Lopez-LPP}, the LPP transition was found by calculating
$\langle S\rangle=\sum_{S} S\Theta(S)$ and determining the parameter values at the threshold of
$\langle S\rangle\sim N$. Here we develop a more systematic approach, in
which we look at the entire distribution $\Theta(S|\tau,a,N)$
in order to explore whether disorder induced LPP exhibits a phase transition, and if
it does, what is the order.
To study the thermodynamic limit, we
find it useful to analyze the fractional mass $\sigma\equiv S/N$, and hence
$\Theta(\sigma|\tau,a,N)$ (or $\Theta(\sigma|\tau,a,L)$ for lattices).
On a square lattice of equal sides $L$ and $N=(L+1)^2$ nodes, we measure
$\Theta(\sigma|\tau,a,L)$ from node $o$ located at
the center of the lattice $(x_o=[[(L+1)/2]],y_o=[[(L+1)/2]])$ with $[[.]]$ indicating the next lowest integer of
the argument.
In Fig.~\ref{ThetaS_vs_N}(a), we show $\Theta(\sigma|\tau,a=10,L)$
with several $L$ and $\tau$. The first interesting feature is the narrow shape of the distribution,
indicating the presence of a characteristic mass $S$ for given $\tau$ and $L$. This suggests focusing on
the most probable value of $\sigma$, labeled $\sigma^*(\tau,a,N)$, i.e.
$\Theta(\sigma^*|\tau,a,N)>\Theta(\sigma|\tau,a,N)$ for all $\sigma\neq\sigma^*$.
Also, we observe that for small $\tau$, as the system size increases, $\sigma^*$
systematically decreases, but in contrast, for large $\tau$, $\sigma^*$ increases.
Between these two cases, we find a $\tau$ for which $\sigma^*(\tau)$ remains virtually constant.
For such $\tau$, labeled $\tau_c^{(\rm Latt)}$, $S\sim N$ since $\sigma^*(\tau_c)=const.$,
signaling the appearance of a phase transition.
From Fig.~\ref{ThetaS_vs_N}(a) we observe that $\sigma^*(\tau_c^{\rm (Latt)})$ is close to
0.20, suggesting a first order transition.
A systematic study of $\sigma^*(\tau,a,N)$ can be carried out with the purpose of understanding in
more detail the phase diagram of the model. In Fig.~\ref{ThetaS_vs_N}(b) we present $\sigma^*$ for $a=10$
and a range of $L$ and $\tau$. The value of $\sigma^*$ was estimated
from $\Theta(\sigma|\tau,a,L)$ by finding an appropriate cubic fit for the peak of the distribution
and calculating the location of its maximum. Three main regimes can be observed for $\tau$
above, close to, and below $\tau_c^{(\rm Latt)}$.
The determination of $\tau_c^{(\rm Latt)}$ as a function of $a$ from simulations is done by inspection
and requires exploring a range of $\tau$ with considerable precision ($\delta\tau\sim10^{-2}$ or
even $10^{-3}$, becoming more sensitive for large $a$) around a certain region, in plots such as Fig.~\ref{ThetaS_vs_N}(b).
For small $\tau$ (say, close to 1 and well below $\tau_c^{(\rm Latt)}$), it is interesting to see that as the
system size increases, $\sigma^*\sim L^{-2}$ indicating that the fraction of the system that is reachable
is smaller than any power of $N$. The situation for $\tau\lesssim\tau_c^{(\rm Latt)}$ is not as clear:
in LPP due to link removal~\cite{Lopez-LPP} there is a regime of power law sizes of $S$, whereas here
such regime is not evident for lattices but seems to be present for ER networks (see below);
our current theory does not shed light on the matter.
For $\tau>\tau_c^{(\rm Latt)}$, we find a progressive increase of $\sigma^*$ with respect to $L$, with
a saturation value that depends on $\tau$; $\sigma^*$ gradually approaches 1
as $\tau\rightarrow\infty$.
To measure $\Theta(\sigma|\tau,a,N)$ in ER networks, we sample over network realizations of $G\in\mathcal{G}$,
and for each $G$ choose an origin $o$ at random.
In Fig.~\ref{ThetaS_vs_N}(c) we present the relevant simulation results.
The qualitative features of $\Theta(\sigma|\tau,a,N)$ for lattices are also present in
ER networks, including the existence of a critical $\tau$, labeled $\tau_c^{(\rm ER)}$.
In contrast to lattices, $\sigma^*(\tau_c^{(\rm ER)})$ is close to 1, indicating a very
dramatic LPP transition, in which a slight change of $\tau$ around $\tau_c^{(\rm ER)}$ leads to a
transition between almost entirely reachable to entirely unreachable global network states.
We also observe that for $\tau<\tau_c^{(\rm ER)}$, there seems to be a power-law decaying relation between
$\sigma^*$ and $N$, consistent with a fractal size object below the threshold, with the decay
exponent $\tau$-dependent.
To analyze the problem further,
we define the quantity $\alpha_{ij}=\ell^{'}_{ij}/\ell_{ij}$, called the length factor,
for each node pair $ij$~\cite{Blunden}, which measures the fractional increase of the path between
$i$ and $j$, and explore the distribution $\Phi(\alpha)$
and its cumulative $F(\alpha)=\int^{\alpha}\Phi(\alpha^{'})d\alpha^{'}$
over realizations of $P_a(w)$ (and $\mathcal{G}$ for ER networks).
Note that the LPP reachability condition is $\alpha_{ij}\leq\tau$.
Figure~\ref{Phialpha-latt}(a) shows the measurement of $F(\alpha|a,L)$ in lattices, and in the inset
the distribution $\Phi(\alpha|a,L)$, which is a well
concentrated function around its maximum $\alpha=\alpha_c$.
The cumulative $F(\alpha)$ increases sharply around $\alpha=\alpha_c$ rapidly
approaching 1, which indicates that many node pairs satisfy $\alpha\leq\alpha_c$.
The sharpness of $\Phi(\alpha)$ increases for larger $L$, and $F(\alpha)$ becomes more step like,
while $\alpha_c$ remains in the same location. Note that $\alpha_c$
can also be determined at the location where $F(\alpha|a,L)$ for increasing $L$ cross over each other.
Increasing $a$, on the other hand, leads to increasing $\alpha_c$, consistent
with path lengths becoming longer under more disordered conditions.
In Fig.~\ref{Phialpha-latt}(c), we focus on $F(\alpha|a,N)$ for ER networks, and observe similar
features to those in the case of lattices (we display the cumulative only as
the small values of path lengths in ER graphs produce large discretization fluctuations),
apart from the asymptotics which appear to be slower.
\subsection{Scaling of path length as a function of disorder}
In order to understand the previous results, we consider the
current knowledge on the problem of optimal paths, which has received
considerable attention in the context of surface growth and domain
walls~\cite{Chen,Huse-Henley,Cieplak} in the physics literature. For the
purpose of clarity we briefly review these results here, starting with
lattices and extending the discussion to networks.
Disordered lattices often exhibit
optimal paths which are self-affine, characterized by
lengths which scale linearly with $\ell_{ij}$,
with a constant prefactor dependent on the roughness exponent related to
the disorder~\cite{Huse-Henley}.
This limit of self-affine paths has become known as the weak
disorder limit.
Another scaling regime has been recognized~\cite{Cieplak} when
the disorder approaches the so-called strong disorder limit.
In this limit, each link weight in the network is very different to any other link weight,
progressively forcing the optimal paths to lie inside the minimum
spanning tree where their lengths scale as $\ell^{d_{\rm opt}}_{ij}$,
$d_{\rm opt}$ being the scaling exponent of the
shortest path in the minimum spanning tree~\cite{Chen,Porto,Ioselevich}.
A general theory explaining these optimal path limits
points out that weak and strong disorder are separated by
a disorder length-scale $\xi$ which depends on the disorder distribution
$P_a(w)$ and some lattice dependent features~\cite{Chen,Wu}.
Optimal paths covering a distance
smaller than $\xi$ are in strong disorder, and those covering a larger distance are in weak disorder
(provided the system is large enough so that $\xi\ll L$).
Thus, $\xi$ is the weak-strong disorder crossover length. The weak and strong disorder
scaling regimes for $\ell^{'}_{ij}$ can be expressed by the scaling relation
\begin{equation}
\ell^{'}_{ij}\sim\xi^{d_{\rm opt}} f\left(\frac{\ell_{ij}}{\xi}\right),
\quad
f(x)=
\left\{
\begin{array}{ll}
x,&x\gg 1(\text{weak})\\
\text{const.},&x\ll 1 (\text{strong}).
\end{array}
\right. \label{lopt}
\end{equation}
The length-scale $\xi$ can be determined by the relation
$\xi=[p_c/(w_c P_a(w_c))]^{\nu}$~\cite{Chen}, where $p_c$ is the percolation
threshold of the lattice, $\nu$ the correlation length exponent of percolation, and
$w_c$ is the solution to the equation $p_c=\int^{w_c}P_a(w)dw$, i.e., the weight
for which the cumulative distribution of $P_a(w)$ is equal to $p_c$.
Based on the previous arguments, we can now postulate the properties of
$\Phi(\alpha)$. We concentrate on the weak disorder limit because it is the only possible
regime in which an LPP transition could take place~\cite{whynotstrong}. In this
regime, based on Eq.~(\ref{lopt}), we find that
$\ell^{'}_{ij}\sim \xi^{d_{\rm opt}}(\ell_{ij}/\xi)=\xi^{d_{\rm opt}-1}\ell_{ij}$,
where the parenthesis corresponds to $\ell_{ij}$ in the scale of the crossover length
$\xi$, and $\xi^{d_{\rm opt}}$ to the length of the path within this crossover length
(see Fig.~3 in Ref.~\cite{Wu}).
This relation indicates that the typical $\alpha$
is given by $\alpha_c\sim\ell^{'}_{ij}/\ell_{ij}=\xi^{d_{\rm opt}-1}\ell_{ij}/\ell_{ij}=\xi^{d_{\rm opt}-1}$.
For $P_a(w)$ of Eq.~(\ref{Pw}), $\xi=(ap_c)^{\nu}$ calculated according to the above formalism, producing
\begin{equation}
\alpha_c^{(\rm Latt)}\sim (ap_c)^{\nu(d_{\rm opt}-1)}.
\label{alphac-latt}
\end{equation}
In order to test this, we present in Fig.~\ref{Phialpha-scaled}(a) (main)
the scaled curves $\Phi(\alpha/\alpha_c)$ for various $L$ and $a$,
where $\alpha_c$ is taken from Eq.~(\ref{alphac-latt}). The collapse is
consistent with our scaling picture, and indicates that indeed there is
a clear path length increase $(ap_c)^{\nu(d_{\rm opt}-1)}$ that explains
the empirical results.
Given the large fraction of node pairs for which $\alpha$ is close to $\alpha_c$,
we postulate that
\begin{equation}
\tau_c^{(\rm Latt)}=\alpha_c^{(\rm Latt)}(a)\sim (ap_c)^{\nu(d_{\rm opt}-1)},
\label{tau-alpha_c}
\end{equation}
i.e., the tolerance necessary to obtain the LPP transition is equivalent to the most probable length factor.
To test this relation, we find by inspection the values of $\tau_c^{\rm(Latt)}$ as a function of $a$, and plot
them in Fig.~\ref{Phialpha-scaled}(a) (inset). The relation between $\tau_c^{\rm(Latt)}$ and $a$ can be fit to
a power law with exponent
$0.297\pm 8$, which is very close to the predicted $\nu(d_{\rm opt}-1)=0.293$, strongly supporting
Eq.~(\ref{tau-alpha_c}). In addition, we find that $\sigma^*(\tau_c^{(\rm Latt)})\sim 0.2\pm0.04$, independent
of $a$.
The previous results indicate that $\sigma^*$ is universal. We have tested this by comparing the shape of
$\sigma^*(\tau,a,L)$ for different $a$, $\tau$, and $L$, and have found that adjusting for a given
combination of these values, the curves for $\sigma^*$ can be made to overlap, supporting universality.
For ER networks, it is known that their path length distributions
are concentrated due to their random structure, leading to a large number of lengths being similar
to an overall typical length~\cite{Braunstein}.
Thus, we simplify our analysis by focusing on the typical lengths
before and after the introduction of disorder. Before disorder sets in, the typical path
length is
\begin{equation}
\ell_{ER}\sim\frac{\log N}{\log\langle k\rangle}.
\end{equation}
However, after weak disorder sets in, it becomes
\begin{equation}
\ell^{'}_{ER}\sim \mu ap_c \log\left(\frac{N^{1/3}}{ap_c}\right)
\end{equation}
which emerges from the relation that is the equivalent to Eq.~(\ref{lopt})
now applied to networks. $\mu$ is a quantity not yet characterized in the literature, and depends on $\langle k\rangle$
and also weakly on $N$ and $a$; for our simulations, $\mu\approx 4$.
The $N$-dependent ratio between $\ell_{ER}$ and $\ell^{'}_{ER}$, which we label $\beta(N)$ (slightly different
to $\alpha$ because the later applies to each pair of nodes, but the former to the overall typical distances),
is given by
\begin{equation}
\beta(N)\sim \mu ap_c\log\langle k\rangle \left[\frac{1}{3}-\frac{\log(ap_c)}{\log N}\right]=
\alpha_{c,\infty}^{\rm(ER)}-\epsilon^{\rm(ER)}(N)
\label{beta}
\end{equation}
where $\alpha_{c,\infty}^{\rm(ER)}\equiv\mu ap_c\log\langle k\rangle/3$, the asymptotic value of $\beta$, and
$\epsilon^{\rm(ER)}(N)\equiv\mu ap_c\log\langle k\rangle\log(ap_c)/\log N$, a finite size correction that vanishes
logarithmically with $N$ as $N\rightarrow\infty$.
In analogy to lattices, we hypothesize that $\tau_c^{\rm(ER)}(N)$ scales as $\beta(N)$, which includes the size
corrections with $N$. Also, in the limit
$N\rightarrow\infty$, we define $\tau_{c,\infty}^{\rm(ER)}\equiv\alpha_{c,\infty}^{\rm(ER)}$.
To test Eq.~(\ref{beta}), we introduce the rescaled variable $\alpha^{'}_{ij}=(\alpha_{ij}+
\epsilon^{\rm(ER)}(N))/\alpha_{c,\infty}^{\rm(ER)}$, and plot
$F(\alpha^{'}|a,N)$ in Fig.~\ref{Phialpha-scaled}(b).
The collapse of the curves is excellent, and supports the validity of our assumptions.
In the thermodynamic limit we expect
\begin{equation}
\tau_{c,\infty}^{(ER)}=\alpha_{c,\infty}^{(ER)}\sim \frac{\mu ap_c\log\langle k\rangle}{3},
\end{equation}
but it is important to keep in mind the large
finite size corrections to $\tau_c^{\rm(ER)}$, which make it more similar in value to Eq.~(\ref{beta})
for finite $N$.
\section{Conclusions}
We study disorder induced Limited Path Percolation
and determine that a percolation transition occurs for a critical tolerance $\tau_c$.
The critical tolerance increases together with the heterogeneity of the disorder.
Numerical results indicate that LPP displays universality. Also, the phase transition is first order,
with the discontinuity in the order parameter $\sigma^*(\tau_c)$ independent of the disorder.
For lattices $\sigma^*(\tau_c)$ is of the order of 0.2; for Erd\H{o}s-R\'enyi networks, our numerical
results suggest it may approach 1 signaling a catastrophic transition.
The concept of length factor applied to the theory of optimal paths predicts
a typical factor $\alpha_c$ which, in turn, predicts $\tau_c$.
We believe the tolerance thresholds predicted here reflect reachability conditions
of some real networks under real failure scenarios (congestion, maintenance, etc.)
better than regular percolation.
\acknowledgments
E.L. was supported by TSB grant SATURN (TS/H001832/1). L.A.B. was supported by UNMdP and FONCyT (PICT/0293/08).
E.L. thanks Felix Reed-Tsochas and Austin Gerig for helpful discussions.
|
\section{Non-invertible economic models. Outline of main results}
Non-invertible dynamical systems have found many applications in various economic models, in which the equilibrium at
time $t+1$ is not uniquely defined by the one at time $t$; instead there may exist several such optimal states at time $t+1$. We refer to these systems as \textit{implicitly defined economic systems}.
In this paper we study the dynamical and ergodic properties of
such systems which present chaotic behavior on certain invariant
sets. Among the economic systems with non-invertible (or
\textit{backward}) dynamics there are the 1-dimensional and the
2-dimensional overlapping generations models, the cash-in-advance
model, the cobweb model with adaptive adjustment and a class of
models representing heterogeneous market agents with adaptively
rational rules. The common feature of all these models is that
they are given by non-invertible dynamical systems and present
chaotic behavior. \ In some of these models, we have
\textit{hyperbolic horseshoes} (as in the cobweb model, see
\cite{O}, \cite{Z}), in others \textit{transversal
homoclinic/heteroclinic orbits from saddle points} (see the
heterogeneous market model, \cite{FG}), or yet in others there exist
\textit{snap-back repellers}, as in the 1-dimensional and
2-dimensional overlapping generations models for certain offer
curves (see \cite{GHT}). Also in the case of unimodal maps
modelling some overlapping generations scenarios,
we have chaotic behavior on \textit{repelling invariant Cantor sets} (as for the logistic map
$F_\nu$ with $\nu>4$, see \cite{Med}, \cite{R}).
For such noninvertible dynamical systems, the inverse limits are
very important since they provide a natural framework in which the
system "unfolds" and they give sequences of intertemporal
equilibria. Also as we will see they are important since many
results from the theory of expansive homeomorphisms can be applied
on inverse limits, in particular those about lifts of invariant
measures. \textit{Equilibrium measures} of Holder potentials are significant examples of invariant measures and they are very important for the evolution of the system.
For instance, the measure of maximal entropy gives the
distribution on the phase space associated to "maximal chaos". The
Sinai-Ruelle-Bowen measure (see \cite{ER}, \cite{Y}) on a
hyperbolic attractor or of an Anosov diffeomorphism is again an
equilibrium measure (for the unstable potential), and gives the
limiting distribution of the forward iterates of Lebesgue-almost
all points in a neighbourhood of the attractor. Thus it is a
\textit{natural measure} or \textit{physical measure} of the
system since it can be actually observed in experiments/computer
simulations.
Another important feature for economic dynamical systems is that
of \textit{stability}. We are interested if a certain model is
\textit{stable} on invariant sets at small fluctuations. In our
case, since we work with infinite sequences of intertemporal
equilibria, one would like to have stability of the shifts on the
inverse limit spaces.
The standard method of studying evolution of a system in economics
is to use random (stochastic) dynamical systems which transfers
exogeneous random "shocks" to the system. However a system which
presents chaotic behavior, has also complicated
\textit{endogeneous} fluctuations.
Also given an implicitly defined economic system with its inverse
limit of intertemporal equilibria and an utility function on these
equilibria, a central goverment/central bank may want to find a
\textit{distribution on the set of intertemporal equilibria} which
maximizes the average value of the utility, but at the same
time keeps the disorder in the system as little as possible in the
long run. If $W(\cdot)$ is a utility function on $\hat \Lambda$
and $\hat \mu$ is a $\hat f$-invariant measure on $\hat \Lambda$
with measure-theoretic entropy $h_{\hat \mu}$, then the maximum in $\hat \mu$ of
the expression $$ \int_{\hat \Lambda} W(\hat x) d\hat \mu(\hat x)
+ h_{\hat \mu} $$ is attained for the \textit{equilibrium measure}
$\hat \mu_W$ of $W$ (see for instance \cite{KH} for the
Variational Principle for Topological Pressure). So the
equilibrium measures may provide a good way to do that, and we
will be able to give geometric and statistical properties of these
measures. One of the defining characteristics of chaos is
sensitive dependence on initial conditions, that is, even if we
start with two initial states that are quite close to each other,
still over time, they may become very far from each other. The
equilibrium measures will permit us to estimate the
\textit{measures of sets of points which stay close} up to $n$
iterations.
We will use the notion of \textbf{chaotic map} several times. We
say that $f$ is \textbf{chaotic} on an invariant set $X$ if $f$ is
topologically transitive on $X$ and $f$ has sensitive dependence
on initial conditions (see for eg. \cite{R}).
\
The \textbf{main sections and results} of the paper are the following:
First we review some important economic models with non-invertible dynamics, like the overlapping generations model,
the cash-in-advance model, the cobweb model with adaptive adjustments and the heterogeneous market model.
A common feature of all these models is the backward dynamics born out of implicitly defined difference equations.
Also in many instances we have chaotic invariant sets for these models, given by horseshoes, or by snap-back repellers,
or by transverse homoclinic orbits. Therefore we have hyperbolicity on certain invariant sets or conjugation of an iterate
with the shift on some 1-sided symbol space $\Sigma_m^+$.
In Theorem \ref{perturbation} we will prove that by slightly
\textbf{perturbing} the parameters of these difference equation, we
obtain again the same dynamical properties, for instance density
of periodic points, topological transitivity, etc.
We study then \textbf{utility functions on inverse limits} for
noninvertible economic systems. Invariant measures for
a dynamical system are very important since they preserve the ergodic and dynamical properties of the system in time;
in fact from any measure one can form canonically an invariant measure by a well-known procedure (see for eg. \cite{KH}).
We will give \textbf{two options to rank utility functions}: \ one
using average values with respect to invariant probability
borelian measures, especially measures of maximal entropy (which
best describe the chaotic distribution of the system over time),
and another by using equilibrium measures of the utility
functions, which give the best average value while keeping the
system as under control as possible.
\textbf{The first option} is given in Theorem \ref{inv} where we rank utility functions of systems given by certain
unimodal maps according to their average values with respect to invariant borelian measures $\hat \mu$ on the inverse limits,
especially with respect to measures of maximal entropy.
For certain expanding systems, namely for logistic maps $F_\nu,
\nu>4$ we are able to compare in Corollary \ref{log} the
\textbf{average utility values} with respect to the corresponding
measures of maximal entropy when perturbing both the discount
factor $\beta$ of the utility $W$, as well as the system parameter $\nu >4$.
Then in Theorem \ref{exp} we will prove that the inverse limits of
certain invariant sets for these models are \textbf{expansive},
and have also the \textbf{specification property}. This will allow
us in Theorem \ref{eq} to show that given a Holder continuous
potential, we can associate to it a special probability measure
called an \textbf{equilibrium measure} (see \cite{KH}, \cite{Bo}
for definitions). This equilibrium measure can be estimated
precisely, on sets of points remaining close to each other up to a
certain positive iterate (i.e on Bowen balls). We can apply these
results to utility functions from economics, which are shown to be
Holder potentials.
\textbf{The second option to rank utility functions} we consider,
is to maximize the \textbf{ratio} between the exponential of the
average value with respect to $\hat \mu$ and the measure $\hat \mu$ of the set of
points from the inverse limit that remain close up to a certain number of iterates. In
this way we find the distribution $\hat \mu$ which maximizes the average
utility value but \textbf{at the same time} keeps the "disorder"
of the system (i.e the entropy of $\hat \mu$) as small as possible
(equivalently the measure of the set of points which shadow $x$ up
to order $n$, is as large as possible). Equilibrium
measures of Holder potentials on the inverse limit have also other
various statistical properties, like \textit{Exponential Decay of
Correlations} on Holder observables (see \cite{Bo}). Then in
Theorem \ref{approx} we approximate the average value of the
utility on inverse limits with those of simpler potentials.
\
Let us remind now several examples of economic dynamical systems,
which are non-invertible:
\textbf{1. The 1-dimensional overlapping generations model.}
This model was proposed initially by Grandmont (\cite{G}) and
studied by various authors (\cite{GHT}, \cite{KS}, \cite{Med}, \cite{Med2}). In
this model we have an economy with constant population divided
into young and old agents, and with a household sector and a
production sector. A typical agent lives for the 2 periods, works
when young and consumes when old and he receives a salary for his
work in the first period. There is a perishable consumption good
and one unit of it is produced with one unit of labour. If money
is supplied in a fixed amount, say $M$, then we have at time $t$,
that $w_t \ell_t = M$, where $w_t$ is the wage rate and $\ell_t$
is the labour. At the same time, $M = p_{t+1} c_{t+1}$ where
$p_{t+1}$ is the expected price of the consumption good at time
$t+1$ and $c_{t+1}$ is the amount of future consumption. Now
agents have an utility function of type $U = V_1(\ell_* - \ell_t)
+ V_2(c_{t+1})$ where $\ell_*$ is the fixed labour endowment of
the young and $\ell_* - \ell_t$ is the leisure at time $t$.
Agents would like to have both as much leisure currently as well
as consumption when old. Thus under the budget constraint from
above $M = w_y \ell_t = p_{t+1} c_{t+1}$ the optimization problem
above gives, by the method of Lagrange multipliers, an implicit
difference equation: \ $\ell_t = \chi(c_{t+1})$, where
$\chi(\cdot)$ is the \textit{offer curve}. Since by assumption one
unit of labour produces one unit of consumption good, we have
$\ell_t = c_t$, hence by denoting $\ell_t$ by $x_t$, we obtain
\begin{equation}\label{1OLG}
y_t = \chi(y_{t+1})
\end{equation}
As Grandmont showed in \cite{G}, in many cases the offer curve is
not given by a monotonic/injective function, making (\ref{1OLG}) a
non-invertible difference equation. Thus for a level of
consumption at time $t$ there may be several levels of optimal
consumption at time $t+1$. In this case we study the
\textit{backward dynamics} of the system, i.e the sequences of future consumption levels allowed by
(\ref{1OLG}).\
The backward dynamics given by relation (\ref{1OLG}) is chaotic in
certain cases. For instance a condition was given by Mitra and
extended in \cite{GHT} in order to guarantee the existence of a
\textit{snap-back repeller}. Let us first recall the definition of a
snap-back repeller (see \cite{Mar}, \cite{Mar2}), and that of the
one-sided shift:
\begin{defn}\label{snap}
Let a smooth function $f:U \to U$, where $U$ is an open set in $\mathbb R^n, n \ge 1$.
Suppose that $p$ is a fixed repelling point of $f$, i.e all the eigenvalues of $Df(p)$ are larger than 1 in absolute
value, and assume that there exists another point $x_0 \ne p$ in a repelling neighbourhood of $p$, so that
$f^m(x_0)= p$ and $\text{det}Df(f^i(x_0)) \ne 0, 1 \le i \le m$. Then $p$ is called a \textit{snap-back
repeller} of $f$.
\end{defn}
\begin{defn}\label{shift}
We will denote by $\Sigma_m^+$ (where $m \ge 2$) the space of
1-sided infinite sequences formed with $m$ symbols, i.e
$\Sigma_m^+ = \{(i_0, i_1, i_2, \ldots), i_j \in \{1, \ldots, m\},
j \ge 0\}$. We have the \textit{shift map} on $\Sigma_m^+$,
namely $\sigma_m: \Sigma_m^+ \to \Sigma_m^+, \ \sigma_m(i_0, i_1,
\ldots) = (i_1, i_2, \ldots)$. The space $\Sigma_m^+$ is compact
with the product topology.
\end{defn}
Snap-back
repellers appear only for non-invertible maps, and are
important since they are similar to transverse homoclinic orbits (see \cite{KH} for eg.)
Marotto proved the following:
\begin{unthm}[Marotto]\label{marotto}
Let $p$ a snap-back repeller for a smooth non-invertible
map $f$ and $\mathcal{O}(x_0)$ a homoclinic orbit of
$x_0$ towards the repelling fixed point $p$, i.e $\mathcal{O}(x_0)
= \{\ldots, x_{-i}, \ldots, x_0, f(x_0), \ldots, p\}$, with
$f(x_{-i}) = x_{-i+1}, i \ge 1$. Then in any neighbourhood of the
orbit $\mathcal{O}(x_0)$ there exists a Cantor set $\Lambda$ on
which some iterate of $f$ is topologically conjugated to the shift
on the space $\Sigma_2^+$ of one-sided infinite sequences on 2
symbols. Hence $f$ itself is chaotic on $\Lambda$.
\end{unthm}
For many economic models, the offer curve $\chi(\cdot)$ is given by a smooth (or piecewise smooth) \textit{unimodal map}
(see \cite{G}, \cite{GHT}, \cite{Med}, \cite{Med2}). We shall recall some of their properties; for more information, see \cite{KS}, \cite{Med}, \cite{Med2}, etc.
A continuous map $f:[a, b] \to [a, b]$ is called \textit{unimodal}
if $f$ is not monotone and there exists a point $c \in (a, b)$ so
that $f(c) \in [a, b]$ and $f$ is increasing on $[a, c)$ and
decreasing on $(c, b]$. Type A unimodal maps are unimodal maps
satisfying $f(a) = a$ and $f(c) < b$. Type B unimodal maps are
those satisfying $f(a) > a$ and $f(b) = a$. Type C maps are of the
form $f:[a, b] \to \mathbb R$ s. t $f$ is not monotone, $f(a)=f(b)
= a$ and $f(c) > b$. Type C maps are not strictly speaking
unimodal as the map $f$ does not take necessarily values inside
the same interval $[a, b]$, but in general they are considered
"unimodal" too. \ In certain cases when the offer curve $\chi$ is
unimodal, one can find snap-back repellers (see \cite{GHT}):
\begin{unpro}
Let $\chi: I \to I$ be a unimodal smooth function on the unit interval, with a maximum point at $x_m$ and a fixed point at $x^*$. If $\chi^3(x_m) < x^*$, then $x^*$ is a snap-back repeller and thus there exists an invariant Cantor set $\Lambda \subset I$ on which an iterate of $\chi$ is topologically conjugate to the shift; so $\chi$ is chaotic and has positive topological entropy.
\end{unpro}
We will need in conjunction with unimodal maps and their inverse limits, the notions of \textit{topological attractor}
and \textit{asymptotically stable attractor}. First given a continuous map $f:X \to X$ on a metric space and a closed
forward invariant set $K \subset X$, we call the \textit{basin of attraction} of $K$ the set
$B(K):= \{y \in X, \omega(y) \subset K\}$, i.e the set of points having all the accumulation points
of their iterates, contained in $K$. Then we say that $K$ is a \textit{topological attractor}, if
$B(K)$ contains a residual set in an open neighbourhood $U$ of $K$ (i.e the complement of $B(K)$ in $U$
is contained in a countable union of nowhere dense subsets) and if there is no closed forward invariant subset
$K' \subset K$ s.t $B(K)$ and $B(K')$ coincide up to a countable union of nowhere dense sets. \
If $K$ is $f$-invariant (i.e $f(K) = K$), it has arbitrarily close neighbourhood $V$ s.t
$f(V) \subset V$ and the basin $B(K)$ is open, then we say that $K$ is an \textit{asymptotically stable attractor}.
\begin{defn}\label{inverse-limit}
Given a continuous map $f: X \to X$ on a metric space $(X, d)$, we form
the \textbf{inverse limit} $(\hat X, \hat f)$, where $\hat X :=
\{\hat x = (x, x_{-1}, x_{-2}, \ldots), f(x_{-i}) = x_{-i+1}, i
\ge 1\}$ and $\hat f:\hat X \to \hat X, f(x, x_{-1}, \ldots) =
(f(x), x, x_{-1}, \ldots), \hat x \in \hat X$. We consider the
topology induced on $\hat X$ from the infinite product of $X$ with
itself. In fact $\hat X$ is a metric space with the metric
$$d(\hat x, \hat y) = \mathop{\sum}\limits_{i \ge 0}
\frac{d(x_{-i}, y_{-i})}{2^i}, \hat x, \hat y \in \hat X$$
\end{defn}
For a $\mathcal{C}^3$ smooth map $f$ on the interval $[a,
b]$, the \textit{Schwarzian derivative} is $Sf(x):=
\frac{f'''(x)}{f'(x)} - \frac 32 (\frac{f''(x)}{f'(x)})^2, x \in
[a, b]$.
We have then, by collecting several results (see \cite{Med}, \cite{R} and references therein) the following:
\begin{unthm}[Attractors in inverse limit spaces of unimodal maps]
\ a) \ Let $f$ be a type A unimodal map on the interval $[0, 1]$, with $Sf <0$ on $[0, 1]$. If $f^2(c) = f(1) >0$ and
$f'(0) >1$, then $\hat 0 = (0, 0, \ldots)$ is an asymptotically stable attractor and a topological attractor for
$\hat f$ and it is the only topological attractor for $\hat f$.
\ b) \ Let $f:[0, 1] \to [0, 1]$ be a unimodal map of type B with $Sf <0$ and assume that
$f$ has a unique fixed point $p \in (c, 1]$ that is repelling for $f$ s.t $f(0) > p$.
Then the point $\hat p = (p, p, \ldots) \in \widehat{[0, 1]}$ is an asymptotically stable attractor and a topological
attractor for $\hat f$ and it is the only topological attractor of $\hat f$ in $\widehat{[0, 1]}$.
\ c) \ Let $f:[0, 1] \to [0, 1]$ be a unimodal map of type B with
$Sf <0$ and with $f(0) < p$, where $p$ is the unique fixed point
in $(c, 1]$. Assume that $f$ has topological attractor $P$ which
is either chaotic or periodic. Then the basin of attraction of $P$
contains a union of $n$ intervals $A_0, \ldots, A_{n-1}$ with
$f^i(A_0) \subset A_i, 1 \le i \le n-1$. Let $\Lambda$ be the set
of points in $[0, 1]$ that are never attracted to $P$. Then
$\Lambda$ is partitioned as $\Lambda_1 \cup \ldots \cup \Lambda_m$
where $\Lambda_j$ is an $f$-invariant transitive Cantor set and
$f|_{\Lambda_j}$ is conjugate to a subshift of finite type. Then
the shift map $\hat f$ has a unique topological attractor namely
$\hat \Lambda_0$.
\ d) \ Consider the Type C logistic map $F_\nu(x) = \nu x (1-x), x
\in [0, 1]$ for $\nu >4$, and let $\Lambda_\nu :=
\mathop{\cap}\limits_{n \ge 0} F_\nu^{-n}([0, 1])$. Then
$\Lambda_\nu$ is $F_\nu$-invariant and $F_\nu$ is topologically
conjugate to the shift on $\Sigma_2^+$. Also $\hat \Lambda_\nu$ is
an asymptotically stable attractor for $\hat F_\nu$.
\end{unthm}
\textbf{2. The 2-dimensional overlapping generations model.}
As in the 1-dimensional model before, we have an economy with two
sectors, a household and a production sector (see \cite{GHT}). The
household sector is the same as before, hence with perfect
foresight, we have for the offer curve $\chi(\cdot)$: \ $\ell_t =
\chi(c_{t+1})$. By comparison with the previous case, output is
now produced both from labour $\ell_t$ supplied at time $t$ by the
household sector, and by \textit{capital stock} $k_{t-1}$ from the
previous period $t-1$, supplied by non-consuming companies which
tend to maximize their profits. The output $y_t$ is the minimum
between $\ell_t$ and $k_{t-1}/a$, where $1/a$ is the productivityy
of the capital. We assume that the capital stock available at the
begining of period $t+1$ is $ k_t = (1-\delta)k_{t-1} + i_t, $
where $0<\delta<1$ is the depreciation rate of the capital and
$i_t$ is the investment, i.e the portion of the output at time $t$
which is invested in the next period. Thus the consumption at time
$t$ is $c_t = y_t - i_t$, and at equilibrium we have $y_t = \ell_t =
\frac{k_{t-1}}{a}$. One obtains then the second order difference
equation: $$ y_t = \chi[a(1-\delta+\frac 1a)y_{t+1}-ay_{t+2}] $$
Hence by substituting $z_t = y_t$ and $w_t = y_{t+1}$ we obtain
the implicitly defined system of equations:
\begin{equation}\label{2OLG}
\left\{ \begin{array}{ll}
z_t = \chi[a(1-\delta+ \frac {1}{a})z_{t+1} -
a w_{t+1}] \\
w_t = z_{t+1}
\end{array}
\right.
\end{equation}
In this model for certain parameter values (see \cite{GHT}), the fixed point $x^*$ is a snap-back repeller,
thus by the results of Marotto (see \cite{Mar}, \cite{Mar2}) in any neighbourhood of the orbit of the snap-back repeller there is an invariant set on which $f$ is chaotic and conjugate to a 1-sided shift.
\textbf{3. Cash-in-advance model.}
The following model can be found in \cite{MR} or \cite{KSY}. In this economy there exists a central government and a representative agent, where the government consumes nothing and sets monetary policy. There exists also a cash good and a credit good, and the agent has a utility function of type
\begin{equation}\label{util}
\mathop{\sum}\limits_{t=0}^\infty \beta^t U(c_{1t}, c_{2t}),
\end{equation}
where $\beta \in (0, 1)$ is the discount factor. The function $U$
takes the form \ $ U(x, y) = \frac{x^{1-\sigma}}{1-\sigma} +
\frac{y^{1-\gamma}}{1-\gamma}, $ with $\sigma>0, \gamma >0$. The
cash good $c_{1t}$ can be bought with money $m_t$, which is
carried over from period $t-1$. The credit good $c_{2t}$ does not
require cash and can be bought on credit. Each period the agent
has an endowment $y$ and $c_{1t} + c_{2t} = y$. We assume also
that the cash good costs the same price $p_t$ as the credit good.
The agent wants to maximize his utility function by a choice of
$\{c_{1t}, c_{2t}, m_{t+1}\}_{t \ge 0}$ subject to constraints: \ $
p_t c_{1t} \le m_t$, and \ $m_{t+1} \le p_t y + (m_t -
p_tc_{1t}) + \theta M_t - p_tc_{2t},$ where $M_t$ is the money
supply controlled by the government for a constant growth, $M_{t+1}
= (1 + \theta) M_t$. Denote by $x_t = m_t/ p_t$ the level
of real money balance. We obtain then an implicitly defined difference equation giving
$x_t$ in terms of $x_{t+1}$ with the help of a non-invertible map $f$, i.e
\begin{equation}\label{CIA}
x_t = f(x_{t+1})
\end{equation}
For certain parameters, it can be shown (see \cite{MR}) that there exists an invariant interval
$[x_l, x_r]$ such that the map $f$ has a periodic cycle of period 3. Hence according to Li-Yorke classic result (see \cite{LY}),
the map $f$ is chaotic on that interval. In fact it can be shown
that there exists an invariant subset of $[x_l, x_r]$ on which
the map is conjugate to a subshift of finite type.
\textbf{4. Cobweb model with adaptive adjustment-hedging.}
In this model (see \cite{O}) the supplier adjusts his production $x_t$ according to the realities of the market while keeping the intention to reach a profit maximum $\tilde x_{t+1}$. It is met for instance in agricultural markets where farmers who plant for example wheat cannot change their crop during the same year/period. This is a hedging rule
$$
x_{t+1} = x_t + \alpha (\tilde x_{t+1} - x_t),
$$
with $\alpha \in (0, 1)$ the speed of adjustment. The aggregate supply from $n$ identical producers is $X_t = nx_t$, and the price is given by $p_t = \frac{c}{Y_t^\beta}$, where $Y_t$ is the demand at period $t$ and $c$ is a fixed parameter. We assume the market clears at each period, i. e $X_t = Y_t$.
Then after a change of variable we obtain the equation
\begin{equation}\label{cobweb}
z_{t+1} = f_{\alpha, \beta}(z_t) = (1-\alpha) z_t + \frac{\alpha}{z_t^\beta}, \ (\alpha, \beta) \in (0, 1) \times (0, \infty)
\end{equation}
This function has a unique fixed point $z = 1$, which is a repeller if $|f_{\alpha, \beta}'(1)| >1$, i.e if $\beta > \frac{2-\alpha}{\alpha}$.
Then Onozaki et. al. (\cite{O}) showed that there exists a number $\bar \beta > \frac{2-\alpha}{\alpha}$ s.t for each $\beta > \bar \beta$, $f_{\cdot, \beta}(\cdot)$ has a hyperbolic horseshoe in the plane.
\textbf{5. A heterogeneous market model.}
We will give only the final formula for this 2-dimensional non-invertible case; more information can be found in \cite{FG}.
One has to study the dynamics of the non-invertible map:
\begin{equation}\label{HM}
\left\{ \begin{array}{ll}
z_{t+1} = z_t[(1-\alpha) - \alpha\frac{b(1-m_t)}{2B}] \\
m_{t+1} = \text{tanh} [\frac{\beta b}{4} \cdot z_t^2 \cdot
(\frac{b(1-m_t)}{B}+1) + \frac {\beta}{2}(C_2 - C_1)]
\end{array}
\right.
\end{equation}
For this model, Foroni and Gardini proved in \cite{FG} that there
are \textit{saddle cycles} with homoclinic or heteroclinic transverse
intersections for certain parameters, which give rise to chaotic
sets (horseshoes) by Smale's Theorem or its variants (see
\cite{R}, \cite{HL}, etc.).
\textbf{Conclusions:}
In the examples above there exist parametrizations in which the
system given implicitly $z_t = f(z_{t+1})$, has some hyperbolic
set $\Lambda$ (in general without critical points) or a set where an iterate is conjugate to a 1-sided shift. The dynamics/ergodic theory in these two cases are very similar. The hyperbolic
case includes also the case with no contracting directions, i.e
the expanding case. The implicit difference equation gives the
\textit{backward dynamics} of the model. We notice that a point
from the inverse limit $\hat \Lambda$ given by $\hat x = (x,
x_{-1}, \ldots)$ represents in fact a sequence of \textit{future
equilibria} which are \textit{allowed by the backward dynamics};
so in the notation $\hat x = (x, x_{-1}, x_{-2}, \ldots)$, we
start from a level of consumption of $x$, then at time 1 we have a
level of consumption $x_{-1}$, then $x_{-2}$ at time 2, and so on.
\section{Metric and ergodic properties on inverse limits of chaotic economic models.}
For the implicitly defined economic models given before, we have seen that there exist invariant
sets on which the function (or one of its iterates) is conjugated to a shift on a symbol space;
this invariant limit set $\Lambda$ is usually obtained from homoclinic/heteroclinic orbits or snap-back repellers and thus we have
a hyperbolic structure on $\Lambda$ (see \cite{R}, \cite{KH}, \cite{Mar2}, etc.)
Hyperbolicity is understood here in the \textit{endomorphism sense},
in which the unstable directions and unstable manifolds depend on whole sequences of consecutive preimages (i.e elements of $\hat \Lambda$), not only on base
points (see \cite{Ru-carte89}, \cite{M-DCDS06} for definitions). We
include in the hyperbolic case also the case of no contracting
directions, i.e the expanding case. For a hyperbolic map $f$ on a compact invariant set $\Lambda$ and a small enough $\delta>0$, we denote by $W^s_\delta(x)$ the local stable manifold at the point $x \in \Lambda$, and by $W^u_\delta(\hat x)$ the local unstable manifold corresponding to the history $\hat x \in \hat \Lambda$. \
Let us prove that in this non-invertible hyperbolic case we
have stability of the inverse limits:
\begin{thm}\label{perturbation}
Let us consider one of the economic models from Section 1, given by a dynamical system $f$ having a hyperbolic invariant set $\Lambda$. Then given any dynamical system $g$ obtained by a small $\mathcal{C}^2$ perturbation of the parameters of $f$, there exists a $g$-invariant set $\Lambda_g$ and a homeomorphism $H: \hat \Lambda \to \hat \Lambda_g$ such that $\hat g \circ H = H \circ \hat f$. Thus the dynamics of $\hat g$ on $\hat \Lambda_g$ is the same as the dynamics of $\hat f$ on $\hat \Lambda$.
\end{thm}
\begin{proof}
From the discussion and references given in Section 1 we see that each model has, for certain parameter choices,
invariant sets obtained from homoclinic or heteroclinic orbits, snap-back repellers or horseshoes (like the cobweb model).
The hyperbolicity is obtained from Smale's Theorem on transverse
homoclinic or heteroclinic intersections (see \cite{R}) or its
non-invertible variant given by Hale and Lin (\cite{HL}). Now let
$U$ be a neighbourhood of $\Lambda$ s.t $\Lambda =
\mathop{\cap}\limits_{n \in \mathbb Z} f^{-n}(U)$.
Then if $g$ is obtained from $f$ by a small $\mathcal{C}^2$ perturbation, we can form the basic set $\Lambda_g = \mathop{\cap}\limits_{n \in \mathbb Z} g^{-n}(U)$. If $f$ is hyperbolic on $\Lambda$, then also $g$ will be hyperbolic on $\Lambda_g$.
The hyperbolicity is understood as for endomorphisms, since $f$ is not necessarily invertible on $\Lambda$ (for instance for $\Lambda$ obtained from a snap back repeller, there are at least two points in $\Lambda$ with $f$-image equal to the fixed repelling point).
Hence from \cite{M-DCDS06} we infer the existence of a
conjugating homeomorphism $H : \hat \Lambda \to \hat \Lambda_g$ between the
inverse limit of $(\Lambda, f)$ and that of $(\Lambda_g, g)$,
which commutes with the lifts $\hat f$ and $\hat g$.
\end{proof}
Notice also that by perturbations and by lifting to the inverse limit, the topological entropy
is not changed, i.e $h_{top}(g|{\Lambda_g}) = h_{top}(\hat g|_{\hat \Lambda_g}) =
h_{top}(f|_\Lambda) = h_{top}(\hat f|_{\hat \Lambda})$. \
We discuss now the notion of \textit{utility function} on the set of intertemporal equilibria (see for eg. \cite{KS}, \cite{S}).
\begin{defn}\label{ut}
Consider a continuous function $f: X \to X$ which is non-invertible on the compact set $X$ contained in $\mathbb R$ or $\mathbb R^2$, and let $\hat X$ be the inverse limit.
A \textbf{utility function} on $\hat X$ is a function $W: \hat X \to \mathbb R$ given
by $$ W(\hat x) = \mathop{\sum}\limits_{i \ge 0} \beta^i
U(x_{-i}), $$ where $\beta \in (0, 1)$ is called the \textit{discount
factor} and,
\ a) in the case $X \subset (0, \infty)$ we have $$U(x) :=
\frac{\text{min}\{1, x\}^{1-\sigma}}{1-\sigma} +
\frac{(2-\text{min}\{1, x\})^{1-\gamma}}{1-\gamma}, \ x \in X, \
\text{with} \ \sigma >0, \gamma>0.$$
\ b) in the case $X \subset (0, 1) \times (0, 1)$, we have $$ U(x, y):=
\frac{x^{1-\sigma}}{1-\sigma} + \frac{y^{1-\gamma}}{1-\gamma}, \
(x, y) \in X, \ \text{with} \ \sigma>0, \gamma >0.$$
\end{defn}
The discount factor in the definition of $W$ expresses the fact that future levels of consumption in intertemporal equilibria become less and less relevant to a representative consumer. \
In economic models with backward dynamics we form as before the set of intertemporal equilibria i.e the inverse limit $\hat \Lambda$, where $f|_\Lambda : \Lambda \to \Lambda$ is the restriction of the dynamical system $f$ to a compact invariant set $\Lambda$. In general $f$ is assumed hyperbolic on $\Lambda$ or conjugated to a subshift of finite type of 1-sided sequences.
The consumers/agents have a utility function $W$ given on $\hat \Lambda$. A central government would like to know the average value of $W$ over $\hat \Lambda$. The question is \textbf{with respect to which measure on $\hat \Lambda$}?
In general one uses probability measures which are preserved by the system (in fact from any arbitrary probability measure we can form an invariant one, according to Krylov-Bogolyubov procedure, see \cite{KH}). \
Now an intertemporal equilibrium $\hat x \in \hat \Lambda$ represents in fact a sequence of future levels of consumption allowed by the implicit difference equations of our economic model. In reality an agent may preffer some open sets of intertemporal equilibria over others, and thus not all equilibria will have the same weight/importance, so it is important to use invariant probability measures $\hat \mu$ on the space $\hat \Lambda$ of intertemporal equilibria.
Also if we denote by $B_n(\hat x, \vp)$ the set of points $\hat y \in \hat \Lambda$ which $\vp$-shadow the orbit of $\hat x$ up to $n$-th iterate (called also a \textit{Bowen ball} in $\hat \Lambda$), we would like to have the measure $\hat \mu$ of $B_n(\hat x, \vp)$ as large as possible. This means we keep the disorder in the system as small as possible, and is equivalent to: as small an entropy $h_{\hat \mu}$ as possible. Indeed it can be shown in general (Brin-Katok Theorem, see \cite{Ma}) that if $\mu$ is an $f$-invariant ergodic measure on a space $X$, then
for $\mu$-almost all $x \in X$, $$h_{\mu} = \mathop{\lim}\limits_{\vp \to 0} \lim \frac 1n \mu(B_n(x, \vp))$$
For instance in the cash-in-advance model (see \cite{KS},
\cite{KSY}, \cite{S}, etc) the government is controlling controls
the money supply on the market by the growth rule $M_{t+1} =
(1+\theta)M_t$, where $\theta>0$ is the growth rate. For each
$\theta$ there exists a different invariant interval
$[x_l(\theta), x_r(\theta)]$ and inverse limit space $\hat
\Lambda(\theta)$. For a utility function $W$ like
in Definition \ref{ut}, economists are interested also in choosing
the appropiate $\theta$ so that the average value $\int_{\hat
\Lambda(\theta)} W d\hat \mu_\theta$ is largest, where $\hat \mu$
is an invariant probability on $\hat \Lambda(\theta)$. In this way
given a certain utility function, we can adjust the money growth rate $\theta$ in such a
way that the average utility value is largest. Many times we want
to study systems from the point of view of the measure of maximal
entropy, which best describes the chaotic nature of the model. Also one can be interested in adjusting the discount
factor $\beta$ of $W$ in order to maximize the average utility value.
We will say below that a compact invariant set $\Lambda$ is \textbf{basic} for $f$ if there exists an open neighbourhood $V$ of $\Lambda$ s.t $\Lambda = \mathop{\cap}\limits_{n \in \mathbb Z} f^n(V)$ and if $f$ is topologically (forward) transitive on $\Lambda$; such a set is also called \textit{locally maximal} (see \cite{KH}). In general the invariant limit sets we have considered in the economic models so far, are basic by construction.
Let us recall the following result about invariant measures on inverse limits (see for instance \cite{Ru-T}); recall that our hyperbolic case includes also the expanding case.
\begin{unthm} [Invariant Measures on Inverse Limits]
Let $f: \Lambda \to \Lambda $ be a continuous topologically transitive map on a compact metric space $\Lambda$ and let $\hat f: \hat \Lambda \to \hat \Lambda$ be its inverse limits. Then there is a bijective correspondence $\mathcal{F}$ between $f$-invariant measures on $\Lambda$ and $\hat f$-invariant measures on $\hat \Lambda$, given by $\mathcal{F}(\hat \mu) = \pi_* (\hat \mu)$ (where $\pi: \hat \Lambda \to \Lambda, \pi(\hat x) = x$ is the canonical projection).
Moreover if in addition $f$ is hyperbolic on the basic set $\Lambda$, then for any Holder continuous potential $\phi$ on $\Lambda$ there exists a unique equilibrium measure $\hat \mu_{\phi\circ \pi}$ of $\phi\circ \pi$ and $\pi_*(\hat \mu_{\phi \circ \pi}) = \mu_\phi$, where $\mu_\phi$ is the equilibrium measure of $\phi$ on $\Lambda$.
\end{unthm}
We give now a formula for the average value of the utility with respect to \textit{any} invariant measure on the inverse limit.
\begin{thm}\label{inv}
Consider a continuous non-invertible map $f$ defined on an open set $V$ in $\mathbb R^2$ or in $\mathbb R$, which has an invariant basic set $\Lambda$. Let also $W(\hat x) = \mathop{\sum}\limits_{i \ge 0} \beta^i U(x_{-i})$ be a utility function on the inverse limit $\hat \Lambda$ as in Definition \ref{ut}. Then for any $\hat f$-invariant borelian measure $\hat \mu$ on $\hat \Lambda$ we have that the average value $$\int_{\hat \Lambda} W d\hat \mu = \frac{1}{1-\beta} \int_\Lambda U d\mu,$$
where $\mu = \pi_*(\hat \mu)$.
If in addition $f$ is hyperbolic on $\Lambda$ and if $\mu_0$ is the unique $f$-invariant measure of maximal entropy on $\Lambda$ and $\hat \mu_0$ is the unique measure of maximal entropy on $\hat \Lambda$, then $\mu_0 = \pi_*(\hat \mu_0)$ and $\int_{\hat \Lambda} W d\hat \mu_0 = \frac{1}{1-\beta} \int_{\Lambda} U d\mu_0$.
\end{thm}
\begin{proof}
If we take the approximating functions $W_n(\hat x) =
\mathop{\sum}\limits_{i=0}^n \beta^i U(x_{-i})$, then $W_n$ converge uniformly towards $W$
since $||W-W_n|| \le C \beta^n, n \ge 1$. Hence
$\int_{\hat\Lambda} W_n d\hat \mu \mathop{\to} \limits_{n \to \infty}
\int_\Lambda W_n d \hat \mu$. Now recall that the measure $\hat
\mu$ is $\hat f$-invariant hence $$\int_{\hat \Lambda} W_n d\hat
\mu = \int_{\hat \Lambda} W_n\circ \hat f^n d\hat \mu = \int_{\hat
\Lambda} U(f^nx) + \beta U(f^{n-1} x) + \ldots + \beta^n U(x)
d\hat \mu$$
But now from the fact that $\mu = \pi_* (\hat \mu)$ we see that
$\int_{\hat \Lambda} g\circ \pi d\hat \mu = \int_{\Lambda} g d \mu$,
if $g$ is any continuous function on $\Lambda$. From the $f$-invariance of $\mu$ we have
$\int_\Lambda U\circ f^i d\mu = \int_\Lambda U d\mu, i \ge 0$; thus in our case $$\int_{\hat
\Lambda} W_n d\hat \mu = \int_\Lambda U(f^n x) + \ldots + \beta^n
U(x) d\mu(x) = (1 + \beta + \ldots + \beta^n) \int_\Lambda U(x)
d\mu(x)$$ So
from the approximation above, we obtain in conclusion that $$
\int_{\hat \Lambda} W d\hat \mu = \frac{1}{1-\beta} \int_\Lambda U
d\mu$$
In particular from the Theorem on Equilibrium Measures above, we
obtain that the unique measure of maximal entropy on $\Lambda$ is
the projection of the unique measure of maximal entropy on $\hat
\Lambda$, i.e $\mu_0 = \pi_*(\hat \mu_0)$ and from the above, $\int_{\hat \Lambda} W d\hat \mu_0 = \frac{1}{1-\beta} \int_{\Lambda} U d\mu_0$.
\end{proof}
If we consider $\mathcal{C}^2$-perturbations $g$ of a hyperbolic endomorphism $f$ on a basic set $\Lambda$ (including the case of a perturbation of an expanding endomorphism on a basic set), then from Theorem \ref{perturbation} we see that there exists a $g$-invariant basic set $\Lambda_g$ s.t $g$ is hyperbolic on $\Lambda_g$ and there exists a conjugating homeomorphism $H: \hat \Lambda \to \hat \Lambda_g$ with $\hat g \circ H = H \circ \hat f$. Then the measure of maximal entropy on $\hat \Lambda_g$, denoted by $\hat \mu_{0, g}$, is obtained as $H_*(\hat \mu_0)$, where $\hat \mu_0$ is the unique measure of maximal entropy on $\Lambda$. \
Thus in general we an calculate the average value of the utility $W$ with respect to the measure of maximal
entropy $\int_{\hat \Lambda_g} W d\hat \mu_{0, g}$ by applying Theorem \ref{inv} and the fact that $\mu_{0, g} = (\pi_g\circ H \circ \hat f)_*(\hat \mu_0)$, i.e
$$\int_{\hat \Lambda_g}W d\hat \mu_0 = \frac{1}{1-\beta} \int_{\Lambda_g} U d (\pi_g\circ H \circ \hat f)_*(\hat \mu_0)$$
The average values of $U$ on $\hat\Lambda_g$
with respect to the corresponding measures of
maximal entropy, are easier to estimate than those on inverse
limits. Economists can use this information to compare average
utility values with respect to the corresponding measures of
maximal entropy for various perturbations, which in reality are
translated by adjustments of the money growth rates.
A case in which this average utility ranking can be applied nicely is for the 1-dimensional overlapping generations economic model in which the backward dynamics is given by a Type C unimodal map (typically the \textit{logistic function} $F_\nu(x) = \nu x(1-x)$ with $\nu > 4$). In this case a central government can choose \textbf{both} the $\nu$ and the $\beta$ which \textbf{maximize the average utility value} over the set of intertemporal equilibria, with respect to the \textbf{measure of maximal entropy} (i.e the invariant measure describing the chaotic distribution over time).
\begin{cor}\label{log}
Let a family of logistic maps given by $F_\nu(x) = \nu x(1-x), x
\in [0, 1]$ with $\nu >4$; then $F_\nu$ has an invariant Cantor
set $\Lambda_\nu$. Consider also a utility function $W_\beta(\hat x) =
\mathop{\sum}\limits_{i \ge 0} \beta^i U(x_{-i})$ with $U(x) :=
\frac{\text{min}\{1, x\}^{1-\sigma}}{1-\sigma} +
\frac{(2-\text{min}\{1, x\})^{1-\gamma}}{1-\gamma}, \ x \in (0,
1), \ \text{for some} \ \sigma
>0, \gamma>0$. Then $$\int_{\hat \Lambda_\nu} W_\beta d\hat \mu_0 =
\frac{1}{1-\beta} \int_{\Sigma_2^+} U\circ h_\nu^{-1} d \mu_{\frac
12, \frac 12},$$ where $\hat \mu_0$ is the measure of maximal
entropy on $\hat\Lambda_\nu$, $\mu_{\frac 12, \frac 12}$ is the
measure of maximal entropy on $\Sigma_2^+$ and $h_\nu :
\Lambda_\nu \to \Sigma_2^+$ is the itinerary map, i.e $h_\nu(x) =
(j_0, j_1, \ldots)$ s.t $F_\nu^k(x) \in I_{j_k}, k \ge 0$ where
$F_\nu^{-1}([0, 1]) = I_1 \cup I_2, \ I_1 \cap I_2 = \emptyset$.
\end{cor}
\begin{proof}
For the logistic map $F_\nu$ with $\nu >4$ it is well known (see
for instance \cite{R}) that $F_\nu$ has an invariant Cantor set
$\Lambda_\nu$. For $\nu> 2+\sqrt{5}$ the map $F_\nu$ is expanding
in the Euclidean metric, and for $4 < \nu \le 2+ \sqrt{5}$, the
map $F_\nu$ is expanding in a modified metric.
Also recall that $F_\nu^{-1}([0, 1]) = I_1 \cup I_2 \subset [0,
1]$ where the subintervals $I_1, I_2$ are disjoint. Then we have
the itinerary map $h_\nu: \Lambda_\nu \to \Sigma_2^+, h(x) = (j_0,
j_1, \ldots)$ given by $F_\nu^k(x) \in I_{j_k}, k \ge 0, x \in
\Lambda_\nu$. It can be noticed that $h_\nu$ is a homeomorphism
which gives the conjugacy between $F_\nu|_{\Lambda_\nu}$ and
$\sigma_2|_{\Sigma_2^+}$.
Now consider the measure of maximal entropy
$\mu_{\frac 12, \frac 12}$ on $\Sigma_2^+$; we know (see for instance \cite{KH}) that $\mu_{\frac 12, \frac 12}$ gives measure
$\frac {1}{2^k}$ to each of the cylinders $\{\hat\omega = (i_0, \ldots, i_{k-1}, j_{k}, \ldots), j_{k}, \ldots
\in \{1, 2\}\}$ when $i_0, \ldots, i_{k-1}$ are fixed, ranging in $\{1, 2\}$.
From the conjugacy above, $h_\nu^{-1}$ transports the measure of maximal entropy
$\mu_{\frac 12, \frac 12}$ on $\Sigma_2^+$ to the measure of maximal entropy $\mu_0$ on $\Lambda_\nu$, i.e
$(h_\nu^{-1})_*(\mu_{\frac 12, \frac 12}) =
\mu_0$.
And from the Theorem on Invariant Measures on Inverse Limits above, we know that
$\mu_0 = \pi_*(\hat \mu_0)$, where $\hat \mu_0$ is the unique measure of maximal entropy on $\hat
\Lambda_\nu$. \
So by applying Theorem \ref{inv} we obtain that $$\int_{\hat
\Lambda_\nu} W_\beta d\hat \mu_0 = \frac{1}{1-\beta}
\int_{\Sigma_2^+} U \circ h_\nu^{-1} d \mu_{\frac 12, \frac 12}$$
\end{proof}
Since we have an expression for the itinerary map
$h_\nu$ not difficult to approximate, and since the
measure $\mu_{\frac 12, \frac 12}$ is relatively easy to work
with, one can use Corollary \ref{log} to find a pair of
parameters $(\nu, \beta)$ maximizing the \textit{average
utility value with respect to the measure of maximal entropy} $$\int_{\hat \Lambda_\nu} W_\beta(\hat x)
d\hat\mu_0(\hat x)$$
\
We will now consider the \textbf{second ranking option} for
utility functions, i.e with respect to their equilibrium measures.
First we give some general topological dynamics definitions and
results.
\begin{defn}\label{expans}
A homeomorphism $f:X \to X$ on a metric space $X$ is called
\textit{expansive} if there exists a positive constant $\delta_0$
s.t if $d(f^i x, f^i y) < \delta_0, i \in \mathbb Z$ then $x = y$.
\end{defn}
The following property is very important for the existence of equilibrium measures of Holder continuous potentials (see \cite{Bo}, \cite{KH}).
\begin{defn}\label{spec}
Let a metric space $X$ and a continuous map $f:X \to X$. A
specification $S = (\tau, P)$ consists of a finite collection
$\tau = \{I_1, \ldots, I_m\}$ of finite intervals $I_i = [a_i,
b_i] \subset \mathbb Z$ and a map $P:T(S)=\mathop{\cup}\limits_{i
=1}^m I_i \to X$ s.t for any $t_1, t_2 \in I_j \in \tau$, we have
$f^{t_2 - t_1}(P(t_1)) = P(t_2)$. The specification $S$ is said to
be $n$-spaced if $a_{i+1} > b_i +n, 1 \le i \le m$ and the minimal
such $n$ is called the spacing of $S$. Let us denote also by $L(S)
= b_m - a_1$. We say that $S$ is $\vp$-shadowed by a point $x \in
X$ if $d(f^n(x), P(n)) < \vp$ for all $n \in T(S)$; if $T(S)$
contains also negative integers, we shadow with iterates of a
preimage of large order of $x$. The map $f$ has the
\textbf{specification property} if for any $\vp>0$ there exists an
$M = M_\vp \in \mathbb N$ s.t any $M$-spaced specification $S$ is
$\vp$-shadowed by a point of $X$ and for any $q \ge M + L(S)$,
there is a period-$q$ orbit $\vp$-shadowing $S$.
\end{defn}
\textbf{Remark.} In the above Definition, if $x$ is the period-$q$ point used in the shadowing and if $a_1 < 0$, then instead of $f^{a_1}(x)$ we can take $f^{kq+a_1}(x)$, for the smallest integer $k \ge 0$ s.t $0 \le kq+a_1 < q$ (as the map is non-invertible); then use forward iterates of this point $f^{kq+a_1}(x)$ in the shadowing of the specification.$\hfill\square$
Let us consider now a continuous map $f:X \to X$ on a metric space $X$ and its inverse limit $(\hat X, \hat f)$, where $\hat X$ is the space of infinite sequences of consecutive preimages and $\hat f: \hat X \to \hat X$ is the shift homeomorphism.
In the sequel we will consider mixing basic sets $\Lambda$, i.e basic sets for the endomorphism $f$ s.t $f$ is topologically mixing on $\Lambda$. In fact from the Spectral Decomposition Theorem (see \cite{Bo}, \cite{KH}), any basic set can be decomposed into a finite partition $\Lambda_1, \ldots, \Lambda_s$ s.t for each $j$ there is some iterate $f^{k_j}$ which leaves $\Lambda_j$ invariant and which is topologically mixing on $\Lambda_j$.
\begin{thm}\label{exp}
Let us consider one of the examples from Section 1 that has a mixing basic set $\Lambda$ on which $f$ is hyperbolic.
Then the shift homeomorphism $\hat f$ is
expansive and has specification property on the inverse limit $\hat \Lambda$.
\end{thm}
\begin{proof}
First of all let us show that $\hat f$ is expansive on $\hat
\Lambda$. Let $\hat x, \hat y \in \hat \Lambda$ s.t $d(\hat f^i
\hat x, \hat f^i \hat y) < \delta$ for all $i \in \mathbb Z$ and
some small $\delta>0$. Now $f$ is hyperbolic as an endomorphism on
$\Lambda$ which from construction is a locally maximal set, i.e
there exists a neighbourhood $U$ of $\Lambda$ s.t $\Lambda =
\mathop{\cap}\limits_{n \in \mathbb Z} f^n(U)$. Then if $d(\hat
f^i \hat x, \hat f^i \hat y) < \delta, i \in \mathbb Z$, it
follows that $d(f^i x, f^i y) < \delta, i \ge 0$, hence $y \in
W^s_\delta(x)$. On the other hand if $d(x_{-i}, y_{-i}) < \delta,
i \ge 0$, for certain prehistories $\hat x, \hat y \in \hat
\Lambda$, it follows that $y \in W^u_\delta(\hat x)$. Now if
$\Lambda$ is a hyperbolic locally maximal set for $f$ it follows
that it has local product structure (see \cite{KH}); thus
$W^s_\delta(x) \cap W^u_\delta(\hat x) = \{x\}$ for $\delta>0$
small enough, so $x = y$. By repeating this argument for all
preimages $x_{-i}$ we obtain that $x_{-i} = y_{-i}, i \ge 0$.
Therefore $\hat x = \hat y$, and $\hat f$ is expansive on $\hat
\Lambda$.
Let us prove now that $\hat f$ has the specification property on $\hat \Lambda$.
We assumed that $f$ is hyperbolic and topologically mixing on $\Lambda$.
Then as in Theorem 18.3.9 of \cite{KH} we can adapt the proof to endomorphisms to show that $f$ has specification property on $\Lambda$.
In order to prove that $\hat f$ has the specification property on $\hat \Lambda$, let us consider a specification
$\hat S$ in $\hat \Lambda$,
$\hat S = (\hat\tau, \hat P)$, where $\hat \tau$ is a collection of finitely many intervals in
$\mathbb Z$ and $\hat P$ is a correspondence between $T(\hat \tau)$ and $\hat \Lambda$.
Assume that $\hat \tau = \{I_1, \ldots, I_m\}$, with $I_i = [a_i, b_i]$ and that
$\hat P(a_i) = \hat \omega^i = (\omega^i, \omega^i_{-1}, \ldots) \in \hat \Lambda, 1 \le i \le m$.
Consider a small $\vp>0$. We will construct now a
specification $S$ in $\Lambda$ with bigger intervals than those of
$\hat S$. Assume that $\text{diam}(\Lambda) \le 1$ and
take $r = r(\vp)$ so large that $\frac{1}{2^{r}} < \vp/2$. Then we
see that if $d(f^j(x_{-r}), f^j(y_{-r})) < \vp/4, 0 \le j \le r$,
then $d(\hat x, \hat y) < \vp$, where $\hat x = (x, x_{-1},
\ldots), \hat y = (y, y_{-1}, \ldots)$. Hence consider the
specification $S$ in $\Lambda$ of the form $(\tau, P)$, where
$\tau = \{[a_1 -r, b_1] \ldots, [a_m -r, b_m]\}$ and $P(a_i-r) =
\omega^i_{-r}, \ldots, P(b_i) = f^{b_i - a_i}(\omega^i), 1 \le i
\le m$. If $a_1 -r<0$ then instead of $f^{a_1-r}(p)$ we take in the shadowing the iterate $f^{kq+a_1-r}(p)$, for the smallest integer $k \ge 0$ s.t $kq+a_1-r \in [0, q)$. For the other points in the orbit of $p$ used for shadowing we take the
positive iterates of $f^{kq+a_1-r}(p)$, i.e $d(\omega^1_{-r+1}, f^{kq+a_1-r+1}(p)) < \vp/4$, etc.
Now assume that the specification $\hat S$ is $(M + r)$-spaced,
where $M=M(\vp/4)$ is the spacing from the specification property
of $f|_\Lambda$ corresponding to $\vp/4$, and where $r = r(\vp)$
is given above. Then from the specification property of $f$ on
$\Lambda$ it follows that for $q \ge M+L(S)= M + L(\hat S) + r$ there is a period-$q$
orbit $\{p, f(p), \ldots, f^{q-1}(p)\}$ which $\vp/4$-shadows $S$.
Then for $r = r(\vp)$ we can take $\hat M(\vp):= M(\vp/4) + r$,
and the orbit of the periodic point of period $q$, $$\hat p= (f^{kq+a_1-r}(p),
f^{kq+a_1-r-1}(p), \ldots, p, \ldots, f^{kq+a_1-r}(p), \ldots) \in \hat \Lambda$$ We know from
the construction of $S$ that the orbit of $f^{kq+a_1-r}(p)$, $\vp/4$-shadows the composite
chain of points $$\{\omega^1_{-r}, \ldots, \omega^1, \ldots
f^{b_1-a_1}(\omega^1)\} \cup \ldots \cup \{\omega^m_{-r}, \ldots,
\omega^m, \ldots, f^{b_m - a_m}(\omega^m)\}$$
Thus we have $d(\omega^1_{-r}, f^{kq+a_1-r}(p))< \vp/4, \ldots, d(\omega^1,
f^{kq+a_1}(p)) < \vp/4, \ldots, d(f^{b_1 - a_1}(\omega^1), f^{kq+b_1}(p)) < \vp/4$ and so on up to the interval $I_m$ where $d(\omega^m_{-r}, f^{kq+a_m-r}(p))< \vp/4, \ldots, d(\omega^m, f^{kq+a_m}(p)) < \vp/4, \ldots, d(f^{b_m - a_m}(\omega^m), f^{kq+b_m}(p)) < \vp/4$.
We want to prove that the orbit of $\hat p$, $\vp$-shadows the
specification $\hat S$. From above we obtain that
$$
\aligned
d(\hat
\omega_i, \hat f^{a_i}(\hat p)) &= d(\omega^i, f^{kq+a_1-r+a_i}(p)) +
\frac{d(\omega^i_{-1}, f^{kq+a_1-r +a_i-1}(p))}{2} + \ldots +
\frac{d(\omega^i_{-r}, f^{kq+a_1-r+a_i-r} (p))}{2^r} + \ldots \\
&< \vp/4 + \vp/8
+ \vp/2^{r+2} + \frac{1}{2^r} < \vp/2 + \vp/2 = \vp,
\endaligned
$$
which
follows from the way we chose $r$ above, i. e such that $\frac{1}{2^r} <
\vp/2$.
\ Then we can similarly prove these inequalities up to order $b_i$ when:
$$
\aligned
d(\hat f^{b_i - a_i} \hat \omega^i, \hat f^{b_i} \hat p) &= d(f^{b_i-a_i}(\omega^i), f^{kq+a_1-r+b_i}(p)) +
\ldots + \frac{d(f^{b_i - a_i}(\omega^i_{-r}), f^{kq+a_1-r+b_i-r} p)}{2^r} + \ldots \\
&< \vp/4 + \vp/8 + \ldots + \vp/2^{r+2} +
\frac{1}{2^r} < \vp
\endaligned
$$
Since the above estimates can be done for all $i = 1, \ldots, m$ we see that
$\hat p$, $\vp$-shadows the specification $\hat S$ if $\hat S$ is $\hat M(\vp):= (M(\vp/4) + 2r)$-spaced.
We notice that the integer $r = r(\vp)$ does not depend on the
specification $\hat S$; in conclusion for any $\vp>0$ we found a
positive integer $\hat M(\vp)$ so that any $\hat M(\vp)$-spaced
specification $\hat S$ in $\hat \Lambda$ is $\vp$-shadowed by a
point in $\hat \Lambda$, and for any $q \ge \hat M(\vp) + L(\hat
S)$ there exists a period-$q$ orbit $\vp$-shadowing $\hat S$.
In conclusion if $f$ has specification property on $\Lambda$, then
also $\hat f$ has specification property on $\hat \Lambda$ which
finishes the proof of the Theorem.
\end{proof}
A representative agent may want to maximize the average value of
his utility function with respect to a $\hat f$-invariant measure
$\hat \mu$ on $\hat \Lambda$ but \textit{at the same time} to have
as much control on the system as possible in the long run. In
other words a possibility is to maximize the following sum giving
the average value plus the control $h_{\hat \mu}$:
\begin{equation}\label{max}
AC(W) (\hat \mu) = \int_{\hat \Lambda} W d\hat \mu + h_{\hat \mu}
\end{equation}
From the Variational Principle for Topological Pressure (see
\cite{KH} for eg.), we know that $AC(W)(\hat \mu)$ is maximized
for a probability measure called the \textbf{equilibrium measure}
of $W$. If $W$ is Holder continuous and $\hat f$ is expansive then
this measure is unique and will be denoted by $\hat \mu_W$. This
measure has important geometric properties and one can precisely
estimate the measure $\hat \mu_W$ of the \textit{Bowen balls}
$B_n(\hat x, \vp):= \{\hat y \in \hat \Lambda, d(\hat f^i \hat y,
\hat f^i \hat x) < \vp, i = 0, \ldots, n-1\}$ (see for eg. \cite{Bo}, \cite{KH}).
In particular when $W$ is constant, the equilibrium measure of $W$
is the measure of maximal entropy. Equilibrium measures appear
also as \textit{Sinai-Ruelle-Bowen measures} in the case of
hyperbolic attractors (see \cite{ER}, \cite{Y}) which give the
limiting distribution of forward trajectories of Lebesgue-almost
all points in a neighborhood of the attractor. In the case of
\textit{non-invertible hyperbolic repellers} equilibrium measures
of stable potentials appear also as \textit{inverse
Sinai-Ruelle-Bowen measures} (see \cite{M-JSP}), i.e invariant
measures describing the limiting distributions of preimages of
large orders, of Lebesgue almost-all points in a neighbourhood of
the non-invertible repeller.
We have the following Theorem giving the measure of a Bowen ball
$B_n(x, \vp)$ in a metric space (see \cite{KH}); by $S_n\phi(y)$
we denote the \textbf{consecutive sum} $\phi(y) + \phi(f(y)) +
\ldots + \phi(f^{n-1}(y))$.
\begin{unthm}[Bowen's Theorem on Equilibrium Measures.]
Let $(X, d)$ be a compact metric space and $f: X \to X$ an expansive homeomorphism with specification property and
$\phi: X \to \mathbb R$ a Holder continuous potential on $X$. Then there exists exactly one equilibrium measure for
$\phi$ and $$\mu_\phi = \mathop{\lim}\limits_{n \to \infty} \frac{1}{\mathop{\sum}\limits_{y \in
\text{Fix}(f^n)}
e^{S_n\phi(y)}} \mathop{\sum}\limits_{y \in \text{Fix}(f^n)} e^{S_n\phi(y)} \delta_y$$
Moreover we can estimate the measure $\mu_\phi$ of Bowen balls by:
\begin{equation}\label{Bw}
A_\vp e^{S_n\phi(y)-nP(\phi)} \le \mu_\phi(B_n(y, \vp)) \le B_\vp e^{S_n\phi(y) - nP(\phi)}, \ y \in X, n \ge 1,
\end{equation}
where $A_\vp, B_\vp>0$ are positive constants depending only on $\vp$, and $P(\phi)$ is a number called the topological pressure of $\phi$.
\end{unthm}
Now we notice that in the examples from Section 1 presenting a hyperbolic set, they are formed from non-critical homoclinic orbits to repelling fixed points or from horseshoes without critical points.
\begin{thm}\label{eq}
Consider one of the economic systems from Section 1 given by a
non-invertible map $f$ that has a hyperbolic mixing basic set
$\Lambda$ containing no critical points of $f$. Let also a utility
function $W$ defined on the inverse limit space $\hat \Lambda$ as
in Definition \ref{ut}. Then there exists a unique equilibrium
measure $\hat \mu_W$ of $W$ on $\hat \Lambda$ and for any $\vp>0$
there are positive constants $A_\vp, B_\vp$ so that for any $ \hat
x \in \hat \Lambda, n \ge 1$, $$A_\vp e^{S_nW(\hat x) - nP(W)} \le
\hat \mu_W(B_n(\hat x, \vp)) \le B_\vp e^{S_nW(\hat x)-
nP(W)}$$
\end{thm}
\begin{proof}
Let us consider the hyperbolic non-invertible map $f$ restricted
to the compact invariant set $\Lambda \subset \mathbb R^2$ having
an inverse limit $\hat \Lambda$, and $W$ as in Definition
\ref{ut} (the same proof works in the 1-dimensional case). The utility function $W$ has an associated discount
factor $\beta \in (0, 1)$.
We will show that $W(\hat x) = \mathop{\sum}\limits_{i \ge 0}
\beta^i U(x_{-i})$ is Holder continuous on the
metric space $\hat \Lambda$. Let us notice first that for the utility functions of
Definition
\ref{ut}, the function $U$ is Holder continuous. So there exists a constant $C>0$ and an exponent
$\gamma\in (0, 1]$ s. t $|U(x)- U(y)| \le C d(x, y)^\gamma, x, y \in \Lambda$, as the set $\Lambda$ is
compact.\
But $W(\hat x) = U(x) + \beta U(x_{-1}) + \beta^2 U(x_{-2}) +
\ldots$, so $|W(\hat x) - W(\hat y)| \le |U(x) - U(y)| + \beta
|U(x_{-1}) - U(y_{-1})| + \beta^2 |U(x_{-2}) - U(y_{-2})| +
\ldots, \ \hat x, \hat y \in \hat \Lambda$. From the Holder
condition for $U$ we obtain that $|U(x_{-i}) - U(y_{-i})| \le C
d(x_{-i}, y_{-i})^\gamma, i \ge 0$. Hence
\begin{equation}\label{W}
|W(\hat x) - W(\hat y)| \le C \cdot [d(x, y)^\gamma + \beta d(x_{-1}, y_{-1})^\gamma + \ldots], \hat x, \hat y \in \hat \Lambda
\end{equation}
Without loss of generality assume that $\text{diam}(\Lambda) = 1$.
Let us take now two close points $\hat x, \hat y \in \hat \Lambda,
d(\hat x, \hat y) < \delta < < 1$. Recall that we have a
hyperbolic structure on $\Lambda$, and denote by $Df_s(x)$ the
restriction of $Df(x)$ to the stable tangent space at $x$. If $x
\ne y$ are close, then we may have some of their preimages of order 1, $x_{-1}$ and
$y_{-1}$ close as well. Denote by $\lambda:= \frac{1}{\inf_\Lambda
|Df_s|}$; then $1 < \lambda < \infty$ since there are no critical
points in $\Lambda$. Assume also that $\gamma>0$ is taken such
that:
\begin{equation}\label{con}
\beta \lambda^\gamma <1
\end{equation}
This is possible if we take $\gamma>0$ small enough, since $\beta
\in (0, 1)$. From the definition of $\lambda$, we know that
$d(x_{-1}, y_{-1}) \le d(x, y) \lambda$ if $x_{-1}, y_{-1}$ are close too. Let us repeat this
procedure with finite sequences of consecutive preimages $x_{-m}, y_{-m}$ until we have $d(x,
y) \lambda^m
>\vp_0$ for some fixed $\vp_0$; i.e $m$ is the first positive
integer satisfying this condition. Then for a choice of $\hat
x, \hat y$ having on the $m$-th positions respectively $x_{-m}, y_{-m}$, we obtain from (\ref{W}): $$|W(\hat x)- W(\hat y)| \le C[d(x,
y)^\gamma + \beta d(x, y)^\gamma \lambda^\gamma + \ldots + \beta^m
d(x, y)^\gamma \lambda^{m\gamma} + \beta^m]$$ We know however that
$m$ is related to $d(x, y)$ and can be expressed in terms of it.
Indeed from the condition on $m$, we have that $m\log \lambda \ge
\log \frac{\vp_0}{d(x, y)}$ and hence $$\beta^m \le C_1 \cdot d(x,
y)^{\rho'},$$ for some constant $\rho' >0$. This together with the
above relation mean that $$|W(\hat x) - W(\hat y)| \le
\frac{C}{1-\beta \lambda^\gamma} d(x, y)^\gamma + C_1 d(x,
y)^{\rho'}$$ So by taking $\rho := \text{min}\{\rho', \gamma\}$ we
obtain that $|W(\hat x) - W(\hat y)| \le C_2 d(x, y)^\rho$. But
$d(\hat x, \hat y) \ge d(x, y)$, therefore we obtain Holder
continuity in this case, namely $|W(\hat x, \hat y) | \le
C_2 d(\hat x, \hat y)^\rho$.
Now assume that $\hat x, \hat y$ are not as above i.e they do not
shadow each other up to order $m$ but instead, for some $1 \le j
\le m$ there is a preimage $y_{-j}$ far from $x_{-j}$, i.e
$d(x_{-j}, y_{-j}) > \vp_0$ (this follows from the fact that there
are no critical points of $f$ in $\Lambda$). Assume that $\kappa$
is the smallest such $j$. Then $$ \aligned |W(\hat x) - W(\hat y)|
& \le C \left[d(x, y)^\gamma + \beta \lambda^\gamma d(x, y) +
\ldots + \beta^\kappa \lambda^{\kappa \gamma} d(x, y)^\gamma +
\beta^\kappa\right] \\
&\le \frac{C}{1- \beta \lambda^\gamma} d(x,
y)^\gamma + C_1 \beta^\kappa,
\endaligned
$$
for some constants $C, C_1 >0$.
Assume first that $d(x, y)^\gamma \le \beta^\kappa$; then $|W(\hat
x) - W(\hat y)| \le C_2 \beta^\kappa$. But $d(\hat x, \hat y) \ge
\frac{d(x_{-\kappa}, y_{-\kappa})}{2^\kappa} \ge
\frac{\vp_0}{2^\kappa}$. Hence there is a sufficiently small
positive constant $\rho$ and a constant $C_3>0$ (both independent of $\hat x, \hat y$) such
that $|W(\hat x) - W(\hat y)| \le C_3 d(\hat x, \hat y)^\rho$. Now
if we have the other case, i.e $d(x, y)^\gamma \ge \beta^\kappa$,
then $$|W(\hat x)- W(\hat y)| \le C_2 d(x, y)^\gamma \le C_2
d(\hat x, \hat y)^\gamma$$
Hence we proved that $W$ is Holder continuous on $\hat \Lambda$,
i.e there are positive constants $C>0, \rho>0$ so that for all
$\hat x, \hat y \in \hat \Lambda$ we have $$|W(\hat x) - W(\hat
y)| \le C d(\hat x, \hat y)^\rho$$
Now we can use Theorem \ref{exp} in order to prove that the
homeomorphism $\hat f$ is expansive and has specification property
on $\hat \Lambda$. Since we showed that $W$ is Holder
continuous on $\hat \Lambda$ it follows that it has a unique
equilibrium measure $\hat \mu_W$ for which we have the estimates
on the measure of Bowen balls from the previous Bowen's Theorem.
Thus for any $\vp>0$ there are positive constants $A_\vp, B_\vp$
so that for $ \hat x \in \hat \Lambda, n \ge 1$, $$A_\vp
e^{S_nW(\hat x) - nP(W)} \le \hat \mu_W(B_n(\hat x, \vp)) \le
B_\vp e^{S_nW(\hat x)- nP(\phi)}$$
\end{proof}
The previous Theorem gives us good estimates for the measure $\hat
\mu_W$ of the set of points whose iterates remain close to the trajectory of a
certain initial condition, up to $n$ consecutive iterates.
We show now that, if we consider the measure of maximal entropy
$\hat \mu_0$ and \textbf{compare} it to the equilibrium measure
$\hat \mu_W$ on $\hat \Lambda$, then the average utility with
respect to $\hat \mu_W$ is bigger than the average utility with
respect to $\hat \mu_0$.
\begin{cor}\label{ut-en}
In the setting of Theorem \ref{eq} consider the measure of maximal
entropy of $\hat f$ on $\hat \Lambda$ and the
equilibrium measure $\hat \mu_W$ of $W$ on $\hat \Lambda$. Then $$
\int_{\hat \Lambda} W d\hat \mu_W \ge \int_{\hat \Lambda} W d\hat \mu_0$$
\end{cor}
\begin{proof}
From the Variational Principle for topological pressure we know
that $\sup \{h_\nu + \int_{\hat \Lambda} W d\nu, \ \nu \ \hat
f-\text{invariant probability on} \ \hat \Lambda\} = P(W) = h_{\hat
\mu_W} + \int_{\hat \Lambda} W d\hat \mu_W$. Hence since $h_{\hat \mu_0} = h_{top}(\hat f)$ we obtain $$ \int_{\hat
\Lambda} W d\hat \mu_0 + h_{top}(\hat f) \le \int_{\hat \Lambda} W
d\hat \mu_W + h_{\hat \mu_W}$$ Then since $h_{top}(\hat f) \ge h_{\hat \mu_W}$
from the Variational Principle for Entropy (see \cite{KH}), we
obtain the conclusion of the Corollary.
\end{proof}
Given the specific form of our utility function, we can
approximate $\hat \mu_W$ with equilibrium states of simpler
functions. Consider $W_n(\hat x) = \mathop{\sum}\limits_{0 \le i
\le n} \beta^i U(x_{-i}), \ \hat x \in \hat \Lambda, \ \text{for} \n \ge 1$. Similarly as in the proof
of Theorem \ref{eq} we can show that $W_n$ is a Holder function on
$\hat \Lambda$, hence it has an equilibrium state $\hat \mu_{W_n}$
on $\hat \Lambda$.
\begin{thm}\label{approx}
In the setting of Theorem \ref{eq}, let a utility function $W$ on $\hat \Lambda$ and the functions $W_n, n \ge 1$ as above. Then the average value of the utility function with respect to $\hat \mu_W$ can be approximated with those of $W_n$, i.e
$$
|\int_{\hat \Lambda} W d\hat \mu_W - \int_{\hat\Lambda}W_n d\hat \mu_{W_n}| \mathop{\to}\limits_{n \to \infty} 0
$$
\end{thm}
\begin{proof}
From Bowen's Theorem applied to equilibrium measures on $\hat \Lambda$ we have that $$\hat
\mu_\phi = \mathop{\lim}\limits_{n \to \infty}
\frac{1}{\mathop{\sum}\limits_{\hat x \in \text{Fix}(\hat f^n)}e^{S_n\phi(\hat x)}}
\mathop{\sum}\limits_{ \hat x \in \text{Fix}(\hat f^n)}
e^{S_n\phi(\hat x)} \delta_{\hat x},$$ for any Holder continuous
potential $\phi$ on $\hat \Lambda$. Hence since $$||W-W_n|| \le
\frac{\beta^n}{1-\beta} \sup_\Lambda |U|,$$ it follows that $n
\cdot |W-W_n|$ converges uniformly to $0$ and thus $\hat \mu_{W_n}
\to \hat \mu_W$ weakly. Hence $$ \aligned |\int W d\hat \mu_W -
\int W_n d\hat \mu_{W_n}| & \le |\int W d\hat \mu_W - \int W d\hat
\mu_{W_n}| + | \int W d\hat \mu_{W_n} - \int W_n d\hat \mu_{W_n}|
\\ &\le |\int W d\hat \mu_W - \int W d\hat \mu_{W_n}| + \frac{\beta^n}{1-\beta}
\cdot \sup_\Lambda |U|,\endaligned$$ since $||W- W_n|| \le
\frac{\beta^n}{1-\beta} \sup_\Lambda |U|$ and since $\hat
\mu_{W_n}$ is a probability measure. So from the weak
convergence of $\hat \mu_{W_n}$ towards $\hat \mu_W$, we obtain
the conclusion of the Theorem.
\end{proof}
\textbf{Acknowledgements:} This paper is suported by the Sectorial
Operational Programme Human Resources Development (SOP HRD),
financed from the European Social Fund and by the Romanian
Government under the contract number SOP HRD/89/1.5/S/62988.
|
\section{Introduction}
Considerable progress has been made in recent years towards understanding the conditions under which a rational morphism $X \dashrightarrow Y$ can exist between smooth projective varieties $X$ and $Y$ over a field. We can mention, for example, M. Rost's degree formulas and their generalisations (see \cite{Rost}, \cite{Merkurjev}, \cite{LevineMorel}). The results obtained have proven to be particularly successful in their application to the study of projective homogeneous varieties over fields. In the case where $X$ and $Y$ are quadrics, several applications of this approach were already known quite some time ago. Indeed, it was already apparent from the foundational works of A. Pfister in the 1960's that the study of rational morphisms between quadrics has many important applications to the algebraic theory of quadratic forms; for example, to the study of the structure of the Witt ring, or to the construction of fields which exhibit certain arithmetic properties pertaining to quadratic forms. On the other hand, the rich structure theory of the Witt ring unearthed by Pfister permitted the study of rational morphisms between quadrics from a entirely algebraic point of view. In the subsequent decades, the algebraic methods were developed and applied intensively to this area of research. Two of the highlights of this approach are the following results, due to D. Hoffmann and O. Izhboldin respectively:
\begin{theorem}[\cite{Hoffmann1}, Theorem 1] Let $X$ and $Y$ be anisotropic projective quadrics over a field $k$ of characteristic different from 2. If there exists $n \geq 1$ such that $\mathrm{dim}(Y) \leq 2^n - 2 < \mathrm{dim}(X)$, then there are no rational morphisms $X \dashrightarrow Y$. \end{theorem}
\begin{theorem}[\cite{Izhboldin}, Theorem 0.2] Let $X$ and $Y$ be anisotropic projective quadrics over a field $k$ of characteristic different from 2 with $\mathrm{dim}(Y) = 2^n - 1 \leq \mathrm{dim}(X)$ for some $n \geq 0$. If there exists a rational morphism $X \dashrightarrow Y$, then there also exists a rational morphism $Y \dashrightarrow X$. \end{theorem}
Of course, phenomena of this sort are more readily explained nowadays in light of the recent advances in the geometric theory (for example, both the above results can be deduced from Rost's degree formula). Perhaps the most general result available concerning rational morphisms between quadrics is due to N. Karpenko and A. Merkurjev. Before we state it, let us recall a couple of standard definitions. If $X$ is an anisotropic projective quadric over a field $k$ of characteristic different from 2, then the \emph{first Witt index} of $X$, denoted $\mathfrak{i}_1(X)$, is the largest positive integer $i$ such that there exists a degree 1 cycle of dimension $i-1$ on the variety $X_{k(X)}$, where $k(X)$ is the function field of $X$. The \emph{Izhboldin dimension} of $X$, denoted $\mathrm{dim}_{Izh}(X)$, is then defined to be the integer $\mathrm{dim}(X) - \mathfrak{i}_1(X) + 1$. The theorem of Karpenko and Merkurjev now says:
\begin{theorem}[\cite{KarpenkoMerkurjev}, Theorem 4.1] Let $X$ and $Y$ be anisotropic projective quadrics over a field $k$ of characteristic different from 2. If there exists a rational morphism $X \dashrightarrow Y$, then
\begin{enumerate} \item[$\mathrm{(1)}$] $\mathrm{dim}_{Izh}(X) \leq \mathrm{dim}_{Izh}(Y)$.
\item[$\mathrm{(2)}$] $\mathrm{dim}_{Izh}(X) = \mathrm{dim}_{Izh}(Y)$ if and only if there exists a rational morphism $Y \dashrightarrow X$. \end{enumerate} \end{theorem}
If one knows something additional about the first Witt indices of the quadrics involved, then one can start to produce more explicit examples. For instance, Theorems 1.1 and 1.2 can both be recovered from Theorem 1.3 modulo the observation that anisotropic quadrics of dimension $2^n-1$ (for some $n \geq 0$) have first Witt index equal to 1. Although the latter fact was originally proved by D. Hoffmann as a corollary of Theorem 1.1, there are several alternative proofs which are completely independent of Theorem 1.1 (for example, see the paper \cite{Karpenko} of N. Karpenko, where all possible values of the first Witt index are determined). \\
All the above statements include the assumption that the characteristic of the base field is different from 2 (and historically, this is the form in which they were first stated). But it turns out that all these results can be extended to allow for fields of characteristic 2, even when the quadrics involved are not smooth over the base field (see the articles \cite{HoffmannLaghribi1}, \cite{HoffmannLaghribi2} and \cite{Totaro1}). This includes the extreme case of \emph{quasilinear} quadrics, which have no smooth points at all (quasilinear quadrics are those quadrics which are defined by the diagonal part of a symmetric bilinear form over a field of characteristic 2). In fact, several problems which remain open for smooth quadrics have recently been settled for quasilinear quadrics. In particular, we want to mention the results of B. Totaro on the birational geometry of quasilinear quadrics. In the article \cite{Totaro1}, Totaro gives a positive answer to the ``Quadratic Zariski Problem'' for quasilinear quadrics: If $X$ and $Y$ are anisotropic quasilinear quadrics of the same dimension over a field such that there exist rational morphisms from $X$ to $Y$ and from $Y$ to $X$, then $X$ and $Y$ are birational. The proof of this result uses the quasilinear analogue of Theorem 1.3 (due to Totaro) together with the following ``Ruledness Theorem'', proved by Totaro in the same article: If $X$ is an anisotropic quasilinear quadric over a field $k$, then $X$ is birational to $X' \times \mathbb{P}_k^{\mathfrak{i}_1(X) - 1}$ for any subquadric $X' \subset X$ of codimension $\mathfrak{i}_1(X) - 1$ (where the integer $\mathfrak{i}_1(X)$ is defined in the same way as it is for smooth quadrics). For non-quasilinear quadrics, the corresponding problems are still wide open, even in the smooth case.
The goal of the present article is to prove analogues of all the above results for the class of so-called quasilinear $p$-hypersurfaces, which generalises the class of quasilinear quadrics. More precisely, a \emph{quasilinear p-hypersurface} is the projective hypersurface defined by a diagonal form of degree $p$ over a field of characteristic $p>0$. Forms of the latter type are called \emph{quasilinear p-forms}. For evident reasons, the geometry of quasilinear hypersurfaces is only interesting over non-perfect fields. Due in part to the absence of any smooth points on these varieties, one does not have access to the general geometric methods which have proved to be very successful in the study of projective homogeneous varieties over fields. On the other hand, there are several unusual aspects of the theory of quasilinear $p$-forms which suggest that a more algebraic approach is feasible. The main result of the paper is Theorem 5.12. This is an analogue of the main result of \cite{KarpenkoMerkurjev}. The latter result, due to N. Karpenko and A. Merkurjev, immediately implies their Theorem 1.3 as a corollary. Our result should also imply the analogue of Theorem 1.3 for arbitrary quasilinear $p$-hypersurfaces, but there is a technical obstruction which we have currently only managed to overcome for quasilinear quadrics and cubics. This issue is discussed in \S 4. Nevertheless, Theorem 5.12 is already sufficient for several interesting applications which hold for all primes $p$. For example, in \S 6 we prove analogues of Theorems 1.1 and 1.2, as well as some results previously obtained for smooth quadrics by A. Vishik. In the final section, we consider the extension of B. Totaro's results on the birational geometry of quasilinear quadrics to the whole class of quasilinear hypersurfaces. We show that Totaro's ``Ruledness Theorem'' for quasilinear quadrics extends to all primes $p$, but due to the technical issue discussed in \S 4, we are only able to present partial results on the analogue of the ``Quadratic Zariski Problem''. Although all these results were previously known in the case $p=2$, the original proofs do not immediately generalise for larger primes $p$, and we are forced to take a different approach. In particular, we provide new proofs of all the above results for the class of quasilinear quadrics.
The motivation for this paper is the article \cite{Hoffmann2} by D. Hoffmann, where an extensive study of quasilinear $p$-forms was carried out from an algebraic point of view. It was shown there that several aspects of the classical theory of quadratic forms over fields carry over in full generality to this new situation. We recall several notions and results from Hoffmann's paper in \S 2 and \S 3.\\
Throughout this article, $F$ will be a non-perfect field of characteristic $p>0$. $\overline{F}$ will denote a fixed algebraic closure of $F$. By a \emph{scheme} we mean a scheme of finite type over a field. By a \emph{variety}, we mean an integral scheme. A scheme will be called \emph{complete} if it is proper over the base field. If $X$ is a scheme defined over a field $k$, the residue field at a point $x \in X$ will be denoted by $k(x)$. If $X$ is a variety, then we will write $k(X)$ for the function field of $X$. If $L/k$ is any field extension, $X_L$ will denote the scheme $X \times_k \mathrm{Spec}\;L$. Finally, we implicitly assume that all morphisms and rational morphisms of schemes are defined relative to the given base field.
\section{Quasilinear $p$-forms}
In this section we introduce quasilinear $p$-forms and discuss some of their basic properties and invariants. Everything in this section can be found in the article \cite{Hoffmann2}. Since the proofs of the statements we need are all very short, we include them for the reader's convenience.
\begin{definition} Let $V$ be a finite dimensional $F$-vector space. A \emph{quasilinear p-form} (we will often simply say \emph{form} for simplicity) on $V$ is a map $\varphi \colon V \rightarrow F$ satisfying
\begin{enumerate} \item[$\mathrm{(1)}$] $\varphi(\lambda v) = \lambda^p \varphi(v)$ for all $\lambda \in F$ and all $v \in V$.
\item[$\mathrm{(2)}$] $\varphi(v+w) = \varphi(v) + \varphi(w)$ for all $v,w \in V$. \end{enumerate} \end{definition}
We will say that $\varphi$ is a \emph{quasilinear p-form over F} if $\varphi$ is a quasilinear $p$-form on some finite dimensional $F$-vector space. In this case, we denote the underlying space by $V_\varphi$. The \emph{dimension} of $\varphi$ is the dimension of $V_\varphi$, denoted $\mathrm{dim}(\varphi)$. A \emph{morphism} $\varphi \rightarrow \psi$ of forms over $F$ will be an $F$-vector space morphism $V_\varphi \rightarrow V_\psi$ which carries $\varphi$ to $\psi$. If $\varphi$ and $\psi$ are isomorphic, we will write $\varphi \simeq \psi$. We will say that $\varphi$ \emph{is proportional to $\psi$} if there exists $\lambda \in F^*$ such that $\varphi \simeq \lambda \psi$, where $\lambda \psi$ is the form on $V_\psi$ defined by $v \mapsto \lambda \psi(v)$. If $\varphi$ is a quasilinear $p$-form over $F$, then a \emph{subform} $\psi$ of $\varphi$ is the restriction of $\varphi$ to a subspace of $V_\varphi$. We write $\psi \subset \varphi$. The \emph{direct sum} $\varphi \oplus \psi$ and \emph{tensor product} $\varphi \otimes \psi$ of forms $\varphi, \psi$ over $F$ are defined in the obvious way. If $L/F$ is a field extension and $\varphi$ a quasilinear $p$-form over $F$, we write $\varphi_L$ for the form over $L$ obtained by the extension of scalars. If $\varphi$ is a quasilinear $p$-form over $F$, a vector $v \in V_\varphi$ is called \emph{isotropic} if $\varphi(v) = 0$. We say that the form $\varphi$ is \emph{isotropic} if $V_\varphi$ contains a nonzero isotropic vector; $\varphi$ is called \emph{anisotropic} otherwise. By condition (2) in the definition, the subset of isotropic vectors in $V_\varphi$ is actually a subspace, and $\varphi$ is isotropic if and only if it has nonzero dimension. This additivity condition also implies that $\varphi$ is ``diagonalized" in \emph{every} basis of $V_\varphi$; that is, of the form $a_1x_1^p + ... + a_nx_n^p$. We will use the notation $\form{a_1,...,a_n}$ to denote the quasilinear $p$-form $a_1x_1^p + ... + a_nx_n^p$ on the $F$-vector space $F^n$ in its standard basis. If $\varphi$ is a quasilinear $p$-form over $F$, and $L/F$ is a field extension, we define the \emph{value set}
\begin{equation*} D_L(\varphi) \mathrel{\mathop:}= \lbrace \varphi_L(v)\;|\;v \in V_\varphi \otimes_F L \rbrace. \end{equation*}
Since $\varphi$ is additive, $D_L(\varphi)$ is actually an $L^p$-vector subspace of $L$. Clearly it is finite dimensional of dimension $\leq \mathrm{dim}(\varphi)$, and we have $D_L(\varphi) = D_L(\varphi_L)$. \\
The classification of quasilinear $p$-forms over $F$ is given by the following statement:
\begin{proposition}[\cite{Hoffmann2}, Proposition 2.6] Let $\varphi$ be a quasilinear $p$-form over $F$ of dimension $n$, and let $a_1,...,a_m$ be a basis for $D_F(\varphi)$ over $F^p$. Then $m \leq n$ and there is an isomorphism of forms over $F$:
\begin{equation*} \varphi \simeq \form{a_1,...,a_m} \oplus \form{\underbrace{0,...,0}_{n-m}}. \end{equation*}
\begin{proof} Let $W \subseteq V_\varphi$ be the subspace of isotropic vectors. For each $1 \leq i \leq m$, let $v_i \in V_\varphi$ be such that $\varphi(v_i) = a_i$, and let $U$ be the subspace of $V_\varphi$ spanned by the $v_i$. Since the $a_i$ are linearly independent over $F^p$, the $v_i$ are linearly independent over $F$. Therefore, in order to prove the statement, it suffices to show that $V_\varphi = U \oplus W$. Clearly $U \cap W = \lbrace 0 \rbrace$. On the other hand, given any $v \in V_\varphi$, we can find $\lambda_i \in F$ such that $\varphi(v) = \sum_{i=1}^m \lambda_i^p a_i$. Then $\varphi(v) = \varphi(\sum_{i=1}^m \lambda_iv_i)$, so that $v - \sum_{i=1}^m \lambda_i v_i$ is an isotropic vector. Therefore $v \in U + W$, and the statement is proved. \end{proof} \end{proposition}
The isomorphism class of a quasilinear $p$-form $\varphi$ over $F$ is therefore determined by two invariants: the dimension of the subspace of isotropic vectors, and the $F^p$-vector space $D_F(\varphi)$. In particular, the theory is essentially vacuous over perfect fields (i.e. when $F = F^p$), which is why we assume $F$ to be non-perfect. In the case of anisotropic forms, we get:
\begin{corollary} Let $\varphi$ and $\psi$ be anisotropic quasilinear $p$-forms over $F$. Then $\psi$ is isomorphic to a subform of $\varphi$ if and only if $D_F(\psi) \subset D_F(\varphi)$. \end{corollary}
If $\varphi$ is an arbitrary quasilinear $p$-form over $F$, it follows from Proposition 2.2. that there is a unique (up to isomorphism) anisotropic subform $\varphi_{an} \subset \varphi$ such that $\varphi = \varphi_{an} \oplus \form{0,...,0}$. The form $\varphi_{an}$ is called the \emph{anisotropic part} of $\varphi$. The integer $\mathfrak{i}_0(\varphi) \mathrel{\mathop:}= \mathrm{dim}(\varphi) - \mathrm{dim}(\varphi_{an})$ is called the \emph{defect index} of $\varphi$. By the proof of Proposition 2.2, $\mathfrak{i}_0(\varphi)$ is nothing else but the dimension of the subspace of isotropic vectors in $V_\varphi$. We write $\varphi \sim \psi$ whenever $\varphi_{an} \simeq \psi_{an}$ for $p$-forms $\varphi$, $\psi$ over $F$. This is obviously an equivalence relation on the set of isomorphism classes of quasilinear $p$-forms over $F$. The following consequence of Proposition 2.2 will be used several times in the sequel:
\begin{lemma} Let $\varphi = \form{a_1,...,a_n}$ be a quasilinear $p$-form over $F$, and let $L/F$ be any field extension. Then there is a subset $\lbrace a_{j_1},...,a_{j_m} \rbrace \subset \lbrace a_1,...,a_n \rbrace$ such that $(\varphi_L)_{an} \simeq \form{a_{j_1},...,a_{j_m}}$.
\begin{proof} Since $D_L(\varphi) = L^p \cdot a_1 + ... + L^p \cdot a_n$, we can find a subset $\lbrace a_{j_1},...,a_{j_m} \rbrace \subset \lbrace a_1,...,a_n \rbrace$ which constitutes a basis of $D_L(\varphi)$ over $L^p$. The statement then follows from Proposition 2.2. \end{proof} \end{lemma}
We now introduce a special class of forms. In the theory of quadratic forms over fields of characteristic $\neq 2$, an important role is played by the class of so-called \emph{Pfister forms} (namely, tensor products of binary forms which represent 1). The set of isomorphism classes of $n$-fold Pfister forms over such a field $k$ is in bijection with the set of pure symbols in the torsion Milnor $K$-group $K_n^M(k)/2K_n^M(k)$. The projective quadric defined by an $n$-fold Pfister form $\pfister{a_1,...,a_n}$ is a \emph{splitting variety} in the sense that for any extension $L/k$, the quadric has an $L$-rational point if and only if the symbol $\lbrace a_1,...,a_n \rbrace$ is divisible by 2 in $K_n^M(L)$. Let us now recall the equal characteristic analogue of the norm residue isomorphism theorem (formerly the Bloch-Kato conjecture), due to S. Bloch, K. Kato and O. Gabber:
\begin{theorem} [\cite{Kato}, \cite{BlochKato}] For any field $F$ of characteristic $p>0$, there are isomorphisms of abelian groups
\begin{eqnarray*} K_n^M(F)/pK_n^M(F) &\xrightarrow{\sim}& \mathrm{ker}(\wp \colon \Omega_F^n \rightarrow \Omega_F^n/d\Omega_F^{n-1})\\
\lbrace a_1,...,a_n \rbrace & \mapsto & \frac{da_1}{a_1} \wedge ... \wedge \frac{da_n}{a_n}. \end{eqnarray*}
\end{theorem}
Here, $(\Omega_F^\bullet, d)$ is the de Rham complex of absolute differential forms over $F$, and $\wp$ is the inverse Cartier operator (see \cite{Kato} for details). It follows from this result that for any $a_1,...,a_n \in F^*$, the pure symbol $\lbrace a_1,...,a_n \rbrace$ is divisible by $p$ in $K_n^M(F)$ if and only if $[F^p(a_1,...,a_n):F^p] <p^n$. Consider the quasilinear $p$-form
\begin{equation*} \pfister{a_1,...,a_n} \mathrel{\mathop:}= \sum_{0 \leq j_1,...,j_n \leq p-1}(\prod_{i=1}^n a_i^{j_i})x_{j_1,...,j_n}^p \end{equation*}
of dimension $p^n$ on the $F$-vector space $F^{p^n}$ in its standard basis. Clearly $\pfister{a_1,...,a_n} \simeq \pfister{a_1} \otimes ... \otimes \pfister{a_n}$. By the definition, we have $D_L(\pfister{a_1,...,a_n}) = L^p(a_1,...,a_n)$ for every field extension $L/F$. It therefore follows from Proposition 2.2 that the form $\pfister{a_1,...,a_n}_L$ is isotropic if and only if $[L^p(a_1,...,a_n):L^p] < p^n$, which in turn holds if and only if $\lbrace a_1,...,a_n \rbrace = 0$ in $K_n^M(L)/pK_n^M(L)$. The (projective) hypersurfaces defined by the forms $\pfister{a_1,...,a_n}$, which are called \emph{quasi-Pfister forms}, may therefore be regarded (in the above sense) as splitting varieties in the equal characteristic setting. Moreover, it turns out that the quasi-Pfister forms exhibit all the same properties as the Pfister quadratic forms in the mixed characteristic: they are precisely those forms which are ``multiplicative''; up to proportionality, the anisotropic quasi-Pfister forms are characterised by the degree to which they split over the generic point of the associated hypersurface; they have the ``roundness'' property, etc. We refer to the article \cite{Hoffmann2} for further details.\\
It will be convenient to record here the following observation concerning the splitting of quasi-Pfister forms over extensions of the base field:
\begin{lemma} Let $\pi = \pfister{a_1,...,a_n}$ be an anisotropic quasi-Pfister form over $F$. Let $L/F$ be a field extension such that $\pi_L$ is isotropic. Then there is a proper subset $\lbrace a_{j_1},...,a_{j_m} \rbrace \subset \lbrace a_1,...,a_n \rbrace$ such that $(\pi_L)_{an} \simeq \pfister{a_{j_1},...,a_{j_m}}$.
\begin{proof} Let $m$ be such that $[L^p(a_1,...,a_n):L^p] = p^m$. Since $\pi_L$ is isotropic, we have $m<n$. Now, we may choose a subset $\lbrace a_{j_1},...,a_{j_m} \rbrace$ of $\lbrace a_1,...,a_n \rbrace$ such that $L^p(a_1,...,a_n) = L^p(a_{j_1},...,a_{j_m})$. Then the quasi-Pfister form $\pfister{a_{j_1},...,a_{j_m}}$ is anisotropic over $L$, and we have
\begin{equation*} D_L((\pi_L)_{an}) = D_L(\pi) = L^p(a_1,...,a_n) = L^p(a_{j_1},...,a_{j_m}) = D_L(\pfister{a_{j_1},...,a_{j_m}}). \end{equation*}
The statement therefore follows from Corollary 2.3. \end{proof} \end{lemma}
To the arbitrary form $\varphi$, we associate a certain anisotropic quasi-Pfister form. Let us first recall the following invariant:
\begin{definition}[\cite{HoffmannLaghribi1}, \cite{Hoffmann2}] Let $\varphi$ be a quasilinear $p$-form over a field $F$, and let $L/F$ be a field extension. The \emph{norm field} of $\varphi$ (over $L$) is the field
\begin{equation*} N_L(\varphi) \mathrel{\mathop:}= L^p(\frac{a}{b}\;|\;a,b \in D_L(\varphi) \cap L^*). \end{equation*} \end{definition}
Note that if two forms $\varphi$ and $\psi$ are proportional over $F$, then $N_L(\varphi) = N_L(\psi)$ for all extensions $L/F$. The norm field invariant was first introduced in the context of quasilinear quadratic forms by D. Hoffmann and A. Laghribi in \cite{HoffmannLaghribi1}. It turns out that this is a birational invariant of quasilinear $p$-hypersurfaces (see Proposition 4.10), and we will exploit this for the proofs of our main results. A more direct description of the norm field is given by the following lemma:
\begin{lemma} Let $\varphi$ denote the quasilinear $p$-form $\form{a_0,...,a_n}$ over $F$. Then
\begin{equation*} N_F(\varphi) = F^p(\frac{a_1}{a_0},...,\frac{a_n}{a_0}) \end{equation*}
\begin{proof} Since the norm field does not change when we multiply $\varphi$ by a scalar, we may assume that $a_0 = 1$. In this case, it is clear that $F^p(a_1,...,a_n) \subseteq N_F(\varphi)$. But the reverse inclusion also holds, since $D_F(\varphi) = F^p + F^p \cdot a_1 + ... + F^p \cdot a_n \subseteq F^p(a_1,...,a_n)$. \end{proof} \end{lemma}
It follows that $N_L(\varphi)$ is finite dimensional over its subfield $L^p$ for any form $\varphi$ over $F$ and any extension $L/F$. Moreover, if $a_1,...,a_m \in F$ are such that $N_F(\varphi) = F^p(a_1,...,a_m)$, then we have $N_L(\varphi) = N_L(\varphi_L) = L^p(a_1,...,a_m)$ for all extensions $L/F$. Note that the dimension of $N_L(\varphi)$ as an $L^p$-vector space is always a power of $p$. One may therefore define:
\begin{definition} The integer $\mathrm{lndeg}_L(\varphi) \mathrel{\mathop:}= \mathrm{log}_p([N_L(\varphi):L^p])$ is called the (\emph{logarithmic}) \emph{norm degree} of $\varphi$ (over $L$). \end{definition}
\begin{remark} In \cite{Hoffmann2}, the \emph{norm degree} of a form $\varphi$ over $F$ is defined to be the integer $\mathrm{ndeg}_F(\varphi) \mathrel{\mathop:}= p^{\mathrm{lndeg}_F(\varphi)}$. For our purposes, it will be more convenient to take the base $p$-logarithm; for example, see Theorem 4.2 for a result of S. Schr\"{o}er which shows that the integer $\mathrm{lndeg}_F(\varphi)$ determines the size of the singular (non-regular) locus of the projective hypersurface $\lbrace \varphi = 0 \rbrace$. \end{remark}
Finally, we define:
\begin{definition}[\cite{HoffmannLaghribi1}, \cite{Hoffmann2}] Let $\varphi$ be a $p$-form over $F$, and let $L/F$ be a field extension. If $\mathrm{lndeg}_L(\varphi) = m$, and $a_1,...,a_m \in L^*$ are such that $N_L(\varphi) = L^p(a_1,...,a_m)$, then the anisotropic quasi-Pfister form $\widehat{\nu}_L(\varphi) \mathrel{\mathop:}= \pfister{a_1,...,a_m}$ is called the \emph{norm form} of $\varphi$ (over $L$). \end{definition}
By Corollary 2.3, this does not depend on the choice of generators $a_i$ up to isomorphism. Note that we have $\widehat{\nu}_L(\varphi) = \widehat{\nu}_L((\varphi_L)_{an})$, and by the proof of Lemma 2.6, both forms are equal to the anisotropic part of $\widehat{\nu}_F(\varphi)$ over $L$. By definition, the dimension of $\widehat{\nu}_L(\varphi)$ is nothing else but $p^{\mathrm{lndeg}_L(\varphi)}$. If $\varphi_L$ is anisotropic, then by Corollary 2.3 and Lemma 2.8, $\varphi_L$ is proportional to a subform of $\widehat{\nu}_L(\varphi)$.
\section{Quasilinear $p$-forms over extensions of the base field}
In this section we record some general observations about the behaviour of quasilinear $p$-forms and their invariants over extensions of the base field. Again, most of the statements here can be found in the article \cite{Hoffmann2}.\\
The invariant $\mathrm{lndeg}_F(\varphi)$ defined above gives a useful necessary (but not sufficient) condition for an anisotropic quasilinear $p$-form to become isotropic over an extension of the base field.
\begin{proposition}[\cite{Hoffmann2}, Proposition 5.2] Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$, and let $L/F$ be a field extension over which $\varphi$ becomes isotropic. Then $\mathrm{lndeg}_L(\varphi) < \mathrm{lndeg}_F(\varphi)$.
\begin{proof} Recall that for any extension $E/F$, the integer $p^{\mathrm{lndeg}_E(\varphi)}$ is equal to the dimension of the anisotropic part of the norm form $\widehat{\nu}_F(\varphi)$ over $E$. Proving the statement therefore amounts to checking that $\widehat{\nu}_F(\varphi)$ becomes isotropic over $L$. But this is evident, because $\varphi$ is proportional to a subform of $\widehat{\nu}_F(\varphi)$. \end{proof} \end{proposition}
\begin{corollary} Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$, and let $a \in F \setminus F^p$. If $\varphi_{F(\sqrt[p]{a})}$ is isotropic, then $a \in N_F(\varphi)$.
\begin{proof} Let $m = \mathrm{lndeg}_F(\varphi)$, and let $a_1,...,a_m \in F^*$ be such that $N_F(\varphi) = F^p(a_1,...,a_m)$. By Proposition 3.1, the field $N_{F(\sqrt[p]{a})}(\varphi) = F(\sqrt[p]{a})^p(a_1,...,a_m) = F^p(a_1,...,a_m,a)$ has dimension $\leq p^{m-1}$ over $F^p(a) = F(\sqrt[p]{a})^p$. It must therefore have dimension $p^m$ over $F^p$, and so $a \in F^p(a_1,...,a_m) = N_F(\varphi)$. \end{proof} \end{corollary}
This allows us to give another characterisation of the norm field for anisotropic forms:
\begin{corollary} Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$. Let $S$ be the set of all $a \in F$ such that $\varphi_{F(\sqrt[p]{a})}$ is isotropic. Then $N_L(\varphi) = L^p(S)$ for all field extensions $L/F$. \begin{proof} By the remarks following Lemma 2.8, it suffices to prove this for $L=F$. Since $N_F(\varphi)$ is invariant under multiplying $\varphi$ by a scalar, we may assume that $\varphi = \form{1,a_1,...,a_n}$ for some $a_i \in F^*$. Then $N_F(\varphi) = F^p(a_1,...,a_n)$ by Lemma 2.8. Since $\varphi_{F(\sqrt[p]{a_i})}$ is clearly isotropic for all $1 \leq i \leq n$, we have $N_F(\varphi) \subset F^p(S)$. The reverse inclusion follows from Corollary 3.2. \end{proof} \end{corollary}
Let $L/F$ be a finitely generated field extension. Recall that $L/F$ is called \emph{separably generated} if $L$ can be realised as a purely transcendental extension of $F$ followed by a separable algebraic extension. If $L/F$ is separably generated, then there are several rather direct ways to see that any anisotropic quasilinear $p$-form over $F$ remains anisotropic over $L$. More generally, we have:
\begin{proposition} Let $L/F$ be any field extension. Then there exists an anisotropic quasilinear $p$-form $\varphi$ over $F$ such that $\varphi_L$ is isotropic if and only if $L/F$ is not separably generated.
\begin{proof} The extension $L/F$ is not separably generated if and only if $L^p/F^p$ is not separably generated. By Proposition 4.1 in Chapter VIII of \cite{Lang} (we refer to the proof rather than the statement itself), the latter is true if and only if $L^p$ and $F$ are not linearly disjoint over $F^p$, which is a precise translation of the condition in our statement. \end{proof} \end{proposition}
In particular, we have:
\begin{corollary}[\cite{Hoffmann2}, Proposition 5.3] Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$, and let $L/F$ be a separably generated field extension. Then $\varphi_L$ is anisotropic. Moreover, $\mathrm{lndeg}_L(\varphi) = \mathrm{lndeg}_F(\varphi)$.\begin{proof} The first assertion is implicit in Proposition 3.4. The second statement follows by applying the first statement to the norm form $\widehat{\nu}_F(\varphi)$. \end{proof} \end{corollary}
\begin{remark} For finite separable extensions, the first part of this statement may be viewed as a replacement for Springer's Theorem from the theory of quadratic forms, which says that an anisotropic quadratic form remains anisotropic over any odd degree extension of the base field. \end{remark}
Thus in order to study the isotropy behaviour of quasilinear $p$-forms over extensions of the base field, we are essentially reduced to considering finite purely inseparable extensions. We conclude this section by collecting a couple of simple facts with this in mind.
\begin{lemma} Let $\varphi$ be a quasilinear $p$-form over $F$, and let $L/F$ be a finite extension of degree $n$. If $\varphi_L$ is isotropic, then $\varphi$ contains a subform of dimension $\leq n$ which becomes isotropic over $L$.
\begin{proof} Let $\mu_1,...,\mu_n$ be a basis for $L$ over $F$, and suppose that $w \in V_\varphi \otimes_F L$ is a nonzero isotropic vector for $\varphi_L$. Then we can write
\begin{equation*} w = v_1 \otimes \mu_1 + ... + v_n \otimes \mu_n \end{equation*}
for some $v_i \in V_\varphi$. The $v_i$ span a nonzero subspace $W \subset V_\varphi$ of dimension $\leq n$, and the restriction $\varphi|_W$ becomes isotropic over $L$. \end{proof} \end{lemma}
\begin{lemma} Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$, and let $L/F$ be a degree $p$ extension. Then
\begin{enumerate} \item[$\mathrm{(1)}$] $\mathrm{dim}(\varphi_L)_{an} \geq \frac{1}{p} \mathrm{dim}(\varphi)$.
\item[$\mathrm{(2)}$] $\mathrm{lndeg}_L(\varphi) \geq \mathrm{lndeg}_F(\varphi) - 1$, with equality if $\varphi_L$ is isotropic. \end{enumerate}
\begin{proof} For any extension $E/F$, the dimension of ($\varphi_{E})_{an}$ is equal to the dimension of the $E^p$-vector space $D_E(\varphi)$ by Proposition 2.2. In order to prove (1), we therefore have to show that $\mathrm{dim}_{L^p}(D_L(\varphi)) \geq \frac{1}{p}\mathrm{dim}_{F^p}(D_F(\varphi))$. But this is obvious, because
\begin{equation*} p \cdot \mathrm{dim}_{L^p}(D_L(\varphi)) = \mathrm{dim}_{F^p}(D_L(\varphi)) \geq \mathrm{dim}_{F^p}(D_F(\varphi)). \end{equation*}
The inequality in statement (2) now follows by applying (1) to the norm form $\widehat{\nu}_F(\varphi)$. In the case where $\varphi_L$ is isotropic, equality holds by Proposition 3.1. \end{proof} \end{lemma}
\section{Quasilinear $p$-hypersurfaces}
In this section we present some basic observations concerning quasilinear $p$-hypersurfaces and rational morphisms between them. We start with the following lemma:
\begin{lemma} Let $F$ be a field of characteristic $p>0$. Let $f = a_0x_0^p + a_1x_1^p + ... + a_nx_n^p \in F[x_0,...,x_n]$ be a polynomial, and assume that $a_0 \neq 0$. Then $f$ is reducible in $F[x_0,...,x_n]$ if and only if $\frac{a_i}{a_0} \in F^p$ for all $i \in [1,n]$.
\begin{proof} We may assume that $a_0 = 1$. If $a_i \in F^p$ for all $i \in [1,n]$, then $f = (\sum_{i=0}^n \sqrt[p]{a_i} x_i)^p$. Conversely, if $f$ is reducible in $F[x_0,...,x_n]$, then it is certainly reducible in $F(x_1,...,x_n)[x_0]$. It follows that $a_1x_1^p + ... + a_nx_n^p$ is a $p^{th}$-power in $F(x_1,...,x_n)$, and this easily implies that $a_i \in F^p$ for all $i \in [1,n]$. \end{proof} \end{lemma}
Now, let $\varphi$ be a quasilinear $p$-form over $F$ of dimension $d+2$. We consider the projective hypersurface $X_\varphi \mathrel{\mathop:}= \mathrm{Proj}(S^\bullet(V_\varphi^*)/(\varphi)) \subset \mathbb{P}(V_\varphi)$ of dimension $d$ over $F$. A scheme of this type will be called a \emph{quasilinear p-hypersurface}. By Lemmas 2.8 and 4.1, the scheme $X_\varphi$ is integral if and only if $\mathrm{lndeg}_F(\varphi) >0$. In particular, $X_\varphi$ is integral whenever $\varphi$ is anisotropic. If $\mathrm{lndeg}_F(\varphi)>0$, then we let $F(\varphi)$ denote the field of rational functions on $X_\varphi$. Note that $F(\varphi)$ may be realised as a purely transcendental extension of $F$ followed by a purely inseparable extension of degree $p$. If $L/F$ is any extension of fields, then $X_{\varphi_L}$ is canonically isomorphic to $(X_\varphi)_L$, and by construction, $X_\varphi$ has an $L$-rational point if and only if $\varphi_L$ is isotropic. In particular, the anisotropic form $\varphi$ becomes isotropic over the field $F(\varphi)$. We will say that $X_\varphi$ is isotropic (resp. anisotropic) if $\varphi$ is isotropic (resp. anisotropic), and we define $\mathfrak{i}_0(X_\varphi) \mathrel{\mathop:}= \mathfrak{i}_0(\varphi)$. By Corollary 3.5, $X_\varphi$ is isotropic if and only if it has a zero cycle of degree 1. Moreover, if $X_\varphi$ is isotropic, then Proposition 2.2 shows that $X_\varphi$ is a cone over $X_{\varphi_{an}}$ with vertex given by the linear subspace of dimension $\mathfrak{i}_0(X_\varphi) - 1$ corresponding to the subspace of isotropic vectors in $V_\varphi$. It follows that for any $i \geq 0$, we have $\mathfrak{i}_0(X_\varphi)> i$ if and only if $X_\varphi$ has a dimension $i$ cycle of degree prime to $p$ (where by \emph{degree} we mean the degree as a cycle on the ambient projective space). \\
A quasilinear $p$-hypersurface $X$ over $F$ is a twisted form of the $p^{th}$ infinitesimal neighbourhood of a hyperplane in some projective space $\mathbb{P}^n_{\overline{F}}$. In particular, the smooth locus of $X$ is empty. Still, since the base field $F$ is assumed to be non-perfect, it is interesting to ask when $X$ is a regular scheme. The following result, due to S. Schr\"{o}er, shows that this is rarely the case:
\begin{theorem}[\cite{Schroer}, Theorem 3.3] Let $\varphi$ be a quasilinear $p$-form over $F$ of dimension $\geq 2$. Then $X_\varphi$ is a regular scheme if and only if $\mathrm{lndeg}_F(\varphi) = \mathrm{dim}(\varphi) - 1$. If $X_\varphi$ is not regular, then the non-regular locus has codimension $\mathrm{lndeg}_F(\varphi)$ in $X_\varphi$. \end{theorem}
The remainder of this section consists of some general observations concerning rational morphisms between quasilinear $p$-hypersurfaces. Since an isotropic quasilinear $p$-hypersurface is a cone over its anisotropic part (and hence stably birational to its anisotropic part), we may restrict our attention to the anisotropic case. Note that if $\varphi$ and $\psi$ are quasilinear $p$-forms over $F$ of dimension $\geq 2$ with $\mathrm{lndeg}_F(\psi) > 0$, then the existence of a rational map $X_\psi \dashrightarrow X_\varphi$ is equivalent to the isotropy of the form $\varphi_{F(\psi)}$. Indeed, given a rational map $X_\psi \dashrightarrow X_\varphi$, the closure of its graph in $X_\psi \times X_\varphi$ pulls back to a rational point on the generic fibre $(X_\varphi)_{F(\psi)}$ of the canonical projection $X_\psi \times X_\varphi \rightarrow X_\psi$. Conversely, any rational point of $(X_\varphi)_{F(\psi)}$ can be viewed as the generic point of a closed subvariety of $X_\psi \times X_\varphi$ birational to $X_\psi$ over $F$; using the other projection $X_\psi \times X_\varphi \rightarrow X_\varphi$, we get a rational morphism $X_\psi \dashrightarrow X_\varphi$. We will switch between the algebraic and geometric terminology where we feel it is appropriate.\\
It will be useful to observe the following simple fact:
\begin{lemma} Let $\varphi$ and $\psi$ be anisotropic quasilinear $p$-forms over $F$. Then
\begin{enumerate} \item[$\mathrm{(1)}$] $\mathrm{dim}(\varphi_{F(\psi)})_{an} \geq \frac{1}{p} \mathrm{dim}(\varphi)$.
\item[$\mathrm{(2)}$] $\mathrm{lndeg}_{F(\psi)}(\varphi) \geq \mathrm{lndeg}_F(\varphi) - 1$, with equality if $\varphi_{F(\psi)}$ is isotropic. \end{enumerate}
\begin{proof} Since $F(\psi)$ can be realised as a purely transcendental extension of $F$ followed by a degree $p$ extension, this follows from Corollary 3.5 and Lemma 3.8. \end{proof} \end{lemma}
Now, a basic question here is the following:
\begin{question} Let $X$, $Y$ and $Z$ be anisotropic quasilinear $p$-hypersurfaces over $F$. Suppose that there exist rational morphisms $X \dashrightarrow Y$ and $Y \dashrightarrow Z$. Does there exist a rational morphism $X \dashrightarrow Z$? \end{question}
One approach to this problem is suggested by the following classical result:
\begin{proposition} Let $X,Y$ and $Z$ be varieties over a field $k$ with $Z$ complete, and suppose that there are rational morphisms $X \dashrightarrow Y$ and $Y \dashrightarrow Z$. If the image of the generic point of $X$ under the map $X \dashrightarrow Y$ is a regular point of $Y$, then there exists a rational morphism $X \dashrightarrow Z$.
\begin{proof} This is a standard application of the valuative criterion of properness. More explicitly, let $y \in Y$ be the image of the generic point of $X$ under the map $X \dashrightarrow Y$. Since $y$ is regular, there is a valuation ring $R$ of the function field $k(Y)$ with residue field $k(y)$. By the valuative criterion of properness, the map $\mathrm{Spec}\;k(Y) \rightarrow Z$ extends to a morphism $\mathrm{Spec}\;R \rightarrow Z$. Passing to the residue field we, get a morphism $\mathrm{Spec}\;k(y) \rightarrow Z$. Finally, composing this with the natural map $\mathrm{Spec}\;k(X) \rightarrow \mathrm{Spec}\;k(y)$ gives a morphism $\mathrm{Spec}\;k(X) \rightarrow Z$, as we wanted. \end{proof} \end{proposition}
\begin{corollary} Let $Y$ be a complete variety over $F$, and let $\varphi$ and $\psi$ be anisotropic quasilinear $p$-forms of dimension $\geq 2$ over $F$ such that $\psi$ is proportional to a subform of $\varphi$. If there exists a rational morphism $X_\varphi \dashrightarrow Y$, then there exists a rational morphism $X_\psi \dashrightarrow Y$.
\begin{proof} It suffices to treat the case where $\psi$ is a codimension 1 subform of $\varphi$. In this case, $X_\varphi$ is regular at the generic point of $X_\psi$, because the latter is an effective Cartier divisor in $X_\varphi$. The statement therefore follows from Proposition 4.5. \end{proof} \end{corollary}
Now observe that the statement of Question 4.4 depends not on the variety $X$, but only on its generic point. Moreover, the function field $F(X)$ can be realised as a purely transcendental extension of $F$ followed by a purely inseparable extension of degree $p$. Replacing the base field $F$ with a suitable purely transcendental extension of it, we therefore reduce to the case where $X$ is just a point (of degree $p$). Using Lemma 3.7 and Corollary 4.6, we can further reduce to the case where $\mathrm{dim}(Y) \leq p-2$. Taking Proposition 4.5 into consideration, we see that Question 4.4 can be settled affirmatively with a positive answer to the following question:
\begin{question} Let $Y$ be an anisotropic quasilinear $p$-hypersurface of dimension $\leq p-2$ over $F$, and let $y \in Y$ be a closed point of degree $p$. Is it true that there exists a regular closed point $z \in Y$ such that $F(z) \cong F(y)$ over $F$? \end{question}
Of course, this is trivially true in the case where $p=2$. Therefore Question 4.4 has a positive answer for quasilinear quadrics, as was well-known previously. For larger primes $p$, it was essentially asked in \cite{Hoffmann2} whether Question 4.7 has a positive answer with the much stronger condition that the point $z$ be the intersection of the hypersurface $Y$ with a line in the ambient projective space (see Question B in \S 5 of that paper). In general, this is not the case, even for $p=3$. To give an explicit example, let $a,b \in F^*$ be such that $[F^3(a,b):F^3] = 3^2 = 9$. Then the anisotropic cubic form $\varphi = \form{1,a+ba^2,b}$ becomes isotropic over the field $F(\sqrt[3]{a})$, but one easily checks that $\varphi$ has no 2-dimensional subforms which become isotropic over the same extension. Nevertheless, the weaker assertion we make here holds in the case where $p=3$:
\begin{proposition} Question 4.7 (and hence Question 4.4) has a positive answer for $p=2$ and $p=3$.
\begin{proof} If $Y$ is just a point, then there is nothing to prove. We may therefore assume that $p=3$ and $\mathrm{dim}(Y) = 1$. Let $\varphi$ be 3-dimensional anisotropic form over $F$ which defines the hypersurface $Y$. By Lemma 2.8, we either have $\mathrm{lndeg}_F(\varphi) = 1$ or $\mathrm{lndeg}_F(\varphi) = 2$. In the latter case, $Y$ is a regular at all of its points by Theorem 4.2, and again there is nothing to prove. We are therefore left with the case where $\mathrm{lndeg}_F(\varphi) = 1$. Since $\varphi$ becomes isotropic over the residue field $F(y)$, it follows from Proposition 3.1 that we have $\mathrm{lndeg}_{F(y)}(\varphi) = 0$. In other words, the subspace of isotropic vectors for $\varphi_{F(y)}$ is 2-dimensional. For dimension reasons, it follows that every 2-dimensional subform of $\varphi$ over $F$ becomes isotropic over $F(y)$. So if $z$ is a closed point of $Y$ defined by any 2-dimensional subform of $\varphi$, we therefore have $F(z) \cong F(y)$ over $F$. Since all such points are regular, the statement is proved. \end{proof} \end{proposition}
Actually, the proof of Proposition 4.8 shows that for \emph{any} prime $p$, Question 4.7 has a positive answer whenever $\mathrm{dim}(Y) \leq 1$. In particular, for the prime 5 we only need to treat the case where $\mathrm{dim}(Y) = 2$ or $\mathrm{dim}(Y) = 3$. Suppose that $\mathrm{dim}(Y) = 2$, and let $\varphi$ be a 4-dimensional anisotropic form defining $Y$. By Lemma 2.8, we have $1 \leq \mathrm{lndeg}_F(\varphi) \leq 3$. Again, the proof of Proposition 4.8 shows that Question 4.7 has a positive answer if $\mathrm{lndeg}_F(\varphi) = 1$ or $\mathrm{lndeg}_F(\varphi) = 3$. Therefore the only interesting case is where $\mathrm{lndeg}_F(\varphi) = 2$. The following example suggests that things are already rather more complicated in this situation:
\begin{example} Let $p=5$, and let $a,b \in F^*$ be such that $[F^5(a,b):F^5] = 5^2 = 25$. Choose a polynomial $g \in F^5[s,t]$ of degree $\leq 4$ in both variables $s,t$ so that the form
\begin{equation*} \varphi = \form{1,a,b,g(a,b)} \end{equation*}
is anisotropic over $F$. By the definition of $\varphi$, we have $\mathrm{lndeg}_F(\varphi) = 2$. Moreover, condition on the coefficients $a,b$ implies that there are derivations $D_a, D_b \colon F \rightarrow F$ satisfying
\begin{enumerate} \item[$\mathrm{(1)}$] $D_a(a) = 1$, $D_a(b) = 0$, and
\item[$\mathrm{(2)}$] $D_b(a) = 0$, $D_b(b) = 1$. \end{enumerate}
For indeterminates $T_0,T_1,T_2,T_3$, we can extend these to derivations $D_a, D_b \colon F[T_0,...,T_3] \rightarrow F[T_0,...,T_3]$ by sending the variables to zero. Now, let us identify our form $\varphi$ with the polynomial $T_0^5 + aT_1^5 + bT_2^5 + g(a,b)T_3^5 \in F[T_0,...,T_3]$. Then the derivatives
\begin{equation*} D_a(\varphi) = T_1^5 + \frac{\partial g}{\partial s}(a,b)T_3^5, \end{equation*}
and
\begin{equation*} D_b(\varphi) = T_2^5 + \frac{\partial g}{\partial t}(a,b)T_3^5 \end{equation*}
necessarily vanish at the non-regular points of the hypersurface $X_\varphi$. A direct calculation then shows that the non-regular locus of $X_\varphi$ consists of just one point, with residue field isomorphic to
\begin{equation*} L = F(\sqrt[5]{\frac{\partial g}{\partial s}(a,b)}, \sqrt[5]{\frac{\partial g}{\partial t}(a,b)}, \sqrt[5]{g(a,b) - a\frac{\partial g}{\partial s}(a,b) - b\frac{\partial g}{\partial t}(a,b)}). \end{equation*}
As far as Question 4.7 is concerned, we are only interested in points of minimal degree (in this case, degree 5), so the first task here is to determine conditions on the polynomial $g$ under which $[L:F] = 5$. One can hope that the restrictions imposed on $g$ are sufficiently strong to force the existence of more than one $L$-rational point on the hypersurface $X_\varphi$; by the proof of Proposition 4.8, this would be enough to provide a positive answer to Question 4.7 for the form $\varphi$. \end{example}
We conclude this section by pointing out the following important consequence of Corollary 4.6. It was first proved by D. Hoffmann and A. Laghribi for quasilinear quadratic forms in \cite{HoffmannLaghribi1}, and was later extended to arbitrary quasilinear $p$-forms by D. Hoffmann in \cite{Hoffmann2}:
\begin{proposition}[\cite{Hoffmann2}, Lemma 7.12] Let $\varphi$ and $\psi$ be anisotropic quasilinear $p$-forms over $F$ of dimension $\geq 2$. If there exists a rational morphism $X_\psi \dashrightarrow X_\varphi$, then $N_F(\psi) \subset N_F(\varphi)$.
\begin{proof} Since any two forms which are proportional have the same norm field, we may assume that $\psi = \form{1,a_1,...,a_n}$ for some $a_i \in F^*$. For each $i \in [1,n]$, let $\tau_i$ denote the binary subform $\form{1,a_i}$ of $\psi$. By Corollary 4.6, there are rational maps $X_{\tau_i} \dashrightarrow X_\varphi$ for all $i$. In other words, $\varphi$ becomes isotropic over all the fields $F(\tau_i) \cong F(\sqrt[p]{a_i})$. By Corollary 3.2, we therefore have $a_i \in N_F(\varphi)$ for all $i$, and hence $N_F(\psi) = F^p(a_1,...,a_n) \subset N_F(\varphi)$. \end{proof} \end{proposition}
This result shows that the norm field and norm degree are birational invariants of quasilinear $p$-hypersurfaces. We will make use of this in the next section.
\section{The Izhboldin dimension and the main theorem}
In this section, we prove the main result of the paper, Theorem 5.12.\\
Let $\varphi$ be an anisotropic quasilinear $p$-form of dimension $\geq 2$ over $F$. In analogy with the theory of quadratic forms, we define the integers
\begin{equation*} \mathfrak{i}_1(X_\varphi) = \mathfrak{i}_1(\varphi) \mathrel{\mathop:}= \mathfrak{i}_0(\varphi_{F(\varphi)}) \end{equation*}
and
\begin{equation*} \mathrm{dim}_{Izh}(X_\varphi) \mathrel{\mathop:}= \mathrm{dim}(X_\varphi) - \mathfrak{i}_1(X_\varphi) + 1. \end{equation*}
\noindent The latter integer will be called the \emph{Izhboldin dimension} of $X_\varphi$.
\begin{example} It follows from the first part of Lemma 4.3 that $\mathfrak{i}_1(\varphi) \leq \mathrm{dim}(\varphi) - \frac{1}{p}\mathrm{dim}(\varphi)$ for any anisotropic form of dimension $\geq 2$. Generally speaking, this bound is sharp. Indeed, if $\pi$ is an anisotropic quasi-Pfister form of dimension $p^n$, then it follows from Lemma 2.6 that $\mathfrak{i}_1(\pi) = p^n - p^{n-1}$. \end{example}
\begin{example} An algebraic variety $X$ is called \emph{incompressible} if every rational morphism from $X$ to itself is dominant. In the theory of quadratic forms, an important result of A. Vishik says that any anisotropic quadric $X$ over a field of characteristic $\neq 2$ with $\mathfrak{i}_1(X) = 1$ is incompressible. This result has a key role to play in the proof of Theorem 1.3. In our setting, the corresponding statement is trivial. In fact, if $X$ is an anisotropic quasilinear $p$-hypersurface over $F$ with $\mathfrak{i}_1(X) = 1$, then $X$ only has one rational point over its function field $F(X)$. In other words, there is only one rational map from $X$ to itself. This map is, of course, the identity. \end{example}
In fact, an anisotropic quasilinear $p$-hypersurface $X$ is incompressible if and only if $\mathfrak{i}_1(X) = 1$, as the following lemma shows:
\begin{lemma} Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$ of dimension $\geq 2$, and let $\psi \subset \varphi$ be a subform of codimension $ \leq \mathfrak{i}_1(\varphi) - 1$. Then the form $\psi_{F(\varphi)}$ is isotropic.
\begin{proof} The point is that the subspace of isotropic vectors for $\varphi_{F(\varphi)}$ must intersect the underlying space of $\psi_{F(\varphi)}$ non-trivially for dimension reasons. \end{proof} \end{lemma}
We are going to show (as a consequence of the main theorem) that it is impossible to find a subform of codimension larger than $\mathfrak{i}_1(\varphi) - 1$ which becomes isotropic over $F(\varphi)$. First we will need a couple of lemmas:
\begin{lemma} Let $\varphi$ be an anisotropic quasilinear $p$-form of dimension $\geq 2$ over $F$, and let $\psi \subset \varphi$ be a subform of codimension 1. Then $\varphi_{F(\varphi)} \sim \psi_{F(\varphi)}$.
\begin{proof} We can write $\varphi = \form{a} \oplus \psi$ for some $a \in F^*$. But $\psi$ represents $a$ over the field $F(\varphi)$, so the statement follows from Proposition 2.2. \end{proof} \end{lemma}
\begin{lemma} Let $\varphi$ be an anisotropic quasilinear $p$-form of dimension $\geq 2$ over $F$. Then there exists a purely transcendental field extension $K/F$ and a subform $\psi \subset \varphi_K$ of codimension $\mathfrak{i}_1(\varphi) -1$ such that $\mathfrak{i}_1(\psi) = 1$.
\begin{proof} We may assume that $\varphi = \form{1, a_1,...,a_n}$ for some $a_i \in F^*$. Let $m<n$ be such that $\mathrm{dim}(\varphi_{F(\varphi)})_{an} = m+1$. Reordering the $a_i$ if necessary, we can assume that $\form{1,a_1,...,a_m}_{F(\varphi)}$ is the anisotropic part of $\varphi_{F(\varphi)}$ by Lemma 2.4. Now, let $T_1,...,T_{n-1}$ be indeterminate variables. Then the function field $F(\varphi)$ is $F$-isomorphic to the field
\begin{equation*} F(T_1,...,T_{n-1})(\sqrt[p]{a_1T_1^p + ... + a_{n-1}T_{n-1}^p + a_n}). \end{equation*}
Let $K = F(T_{m+1},...,T_{n-1})$, and consider the subform
\begin{equation*} \psi = \form{1,a_1,...,a_m,a_{m+1}T_{m+1}^p + ... + a_{n-1}T_{n-1}^p + a_n} \end{equation*}
of $\varphi_K$. Then we have
\begin{equation*} \psi_{K(\psi)} \sim \form{1,a_1,...,a_m}_{K(\psi)} \end{equation*}
by Lemma 5.4. But the field $K(\psi)$ is $F$-isomorphic to $F(\varphi)$ by construction, so the form $\form{1,a_1,...,a_m}_{K(\psi)}$ is anisotropic. It follows that $\mathfrak{i}_1(\psi) = 1$. \end{proof} \end{lemma}
\begin{remark} We will show later (see Proposition 6.1) that $\mathfrak{i}_1(\psi) = 1$ for any subform $\psi \subset \varphi$ of codimension $\mathfrak{i}_1(\varphi) -1$ over the base field $F$. The example constructed above (modulo passing to a purely transcendental extension of $F$) will be sufficient for our more immediate concerns. \end{remark}
In the theory of quadratic forms, Theorem 1.3 is actually deduced as a consequence of the following stronger statement. It was first proved over fields of characteristic different from 2 by N. Karpenko and A. Merkurjev. It was extended to smooth quadrics in characteristic 2 in the book \cite{EKM} by R. Elman, N. Karpenko and A. Merkurjev, and to arbitrary quadrics (smooth or otherwise) by B. Totaro.
\begin{theorem}[\cite{KarpenkoMerkurjev} Theorem 3.1, \cite{Totaro1} Theorem 5.1] Let $X$ be an anisotropic quadric over a field $k$, and let $Y$ be a complete variety over $k$ which has no closed points of odd degree. Suppose that $Y$ has a closed point of odd degree over $k(X)$. Then
\begin{enumerate} \item[$\mathrm{(1)}$] $\mathrm{dim}_{Izh}(X) \leq \mathrm{dim}(Y)$.
\item[$\mathrm{(2)}$] If $\mathrm{dim}_{Izh}(X) = \mathrm{dim}(Y)$, then there exists a rational morphism $Y \dashrightarrow X$. \end{enumerate} \end{theorem}
\begin{remark} Let $X$ be an anisotropic quadric satisfying $\mathfrak{i}_1(X)=1$. In the terminology of \cite{Haution} \S 10, Theorem 5.7 says that $X$ is \emph{strongly 2-incompressible}. For an arbitrary variety $X$, the degree to which $X$ fails to satisfy the weaker notion of \emph{incompressibility} (see Example 5.2) can be measured by the minimum dimension of the image of a rational morphism from $X$ to itself. In the case where the variety $X$ is regular, this integer is commonly referred to as the \emph{canonical dimension} of $X$. The first part of Theorem 5.7 implies that the canonical dimension of a smooth anisotropic quadric is equal to the Izhboldin dimension $\mathrm{dim}_{Izh}(X)$. This includes the previously mentioned result which says that an anisotropic quadric $X$ is incompressible if and only if $\mathfrak{i}_1(X) = 1$. \end{remark}
We will now prove an analogue of Theorem 5.7 for quasilinear $p$-hypersurfaces. The approach given here is rather different from the one for quadrics given in \cite{KarpenkoMerkurjev} and \cite{Totaro1}. In fact, the latter approach does not work for quasilinear $p$-hypersurfaces whenever $p>2$. Indeed, the argument given in \cite{KarpenkoMerkurjev} and \cite{Totaro1} makes essential use of the fact that the Chow group of zero cycles on a projective quadric injects to the integers via the degree map. This is also true for isotropic quasilinear $p$-hypersurfaces, but is in general false in the anisotropic case (which is precisely the case needed for the proof). For example, if $p>2$, and $X$ is a quasilinear $p$-hypersurface of dimension 1 (i.e. a curve), then one can show that $\mathrm{CH}_0(X)$ contains nonzero $p$-torsion if and only if $X$ is regular. The proof which we present here makes no use of intersection theory, but rather exploits properties of the norm field and norm degree invariants introduced earlier. The important observation is the following lemma:
\begin{lemma} Let $\varphi$ and $\psi$ be anisotropic quasilinear $p$-forms of dimension $\geq 2$ over $F$, and let $L$ be a field such that $F \subset L \subset F(\psi)$. If the form $\varphi_{F(\psi)}$ is isotropic, then $\varphi_L$ is isotropic if and only if $\mathrm{lndeg}_L(\varphi) < \mathrm{lndeg}_F(\varphi)$.
\begin{proof} One direction was already proved in Proposition 3.1. The interesting part is the converse. So suppose that $\mathrm{lndeg}_L(\varphi) < \mathrm{lndeg}_F(\varphi)$, and suppose for the sake of contradiction that $\varphi_L$ is anisotropic. Then since $\varphi_{F(\psi)}$ is isotropic, we have $\mathrm{lndeg}_{F(\psi)}(\varphi) < \mathrm{lndeg}_L(\varphi)$ by Proposition 3.1. But this implies that $\mathrm{lndeg}_{F(\psi)}(\varphi) \leq \mathrm{lndeg}_F(\varphi) - 2$, which is impossible by the second part of Lemma 4.3. \end{proof} \end{lemma}
\begin{proposition} Let $X$ and $Y$ be anisotropic quasilinear $p$-hypersurfaces over $F$. Suppose that there exists a rational morphism $X \dashrightarrow Y$, and let $Z$ denote (the closure of) its image in $Y$. Then there exists a rational morphism $Z \dashrightarrow X$.
\begin{proof} Let $\psi$ and $\varphi$ be anisotropic forms over $F$ defining $X$ and $Y$ respectively. We need to show that $\psi$ becomes isotropic over the field $F(Z)$. But we have a natural embedding $F(Z) \subset F(\psi)$ over $F$, and since the form $\psi_{F(\psi)}$ is isotropic, Lemma 5.9 shows that it is sufficient to check that $\mathrm{lndeg}_{F(Z)}(\psi) < \mathrm{lndeg}_F(\psi)$. Now, since there exists a rational map $X \dashrightarrow Y$, we have an inclusion $N_E(\psi) \subset N_E(\varphi)$ for all extensions $E/F$ by Proposition 4.10. Suppose that $\mathrm{lndeg}_{F(Z)}(\psi) = \mathrm{lndeg}_F(\psi)$. Since $Z$ is a subvariety of $Y$, the form $\varphi_{F(Z)}$ is isotropic, and hence
\begin{equation} \label{eq1} \mathrm{lndeg}_{F(Z)}(\varphi) < \mathrm{lndeg}_F(\varphi) \end{equation}
by Proposition 3.1. By part (2) of Lemma 4.3, we also have
\begin{equation} \label{eq2} \mathrm{lndeg}_{F(\psi)}(\psi) = \mathrm{lndeg}_F(\psi) - 1 = \mathrm{lndeg}_{F(Z)}(\psi) - 1. \end{equation}
Now, since $N_{F(Z)}(\psi) \subset N_{F(Z)}(\varphi)$, it follows from \eqref{eq2} that $\mathrm{lndeg}_{F(\psi)}(\varphi) \leq \mathrm{lndeg}_{F(Z)}(\varphi) - 1$. But then combining this with \eqref{eq1}, we get $\mathrm{lndeg}_{F(\psi)}(\varphi) \leq \mathrm{lndeg}_F(\varphi) - 2$, which is impossible by the second part of Lemma 4.3. Hence $\mathrm{lndeg}_{F(Z)}(\psi) < \mathrm{lndeg}_F(\psi)$, and the proof is complete. \end{proof} \end{proposition}
\begin{corollary} Let $f \colon X \dashrightarrow Y$ be a rational morphism of anisotropic quasilinear $p$-hypersurfaces over $F$, and let $Z$ denote (the closure of) its image in $Y$. If $\mathfrak{i}_1(X) = 1$, then $X$ and $Z$ are birational via $f$.
\begin{proof} There exists a rational morphism $Z \dashrightarrow X$ by Proposition 5.10, and since $\mathfrak{i}_1(X) = 1$, the composition $X \dashrightarrow Z \dashrightarrow X$ is the identity (see Example 5.2). \end{proof} \end{corollary}
A point $y$ on a variety $Y$ over a field $k$ will be called \emph{separable} if the residue field extension $k(y)/k$ is separably generated. Here is our version of Theorem 5.7:
\begin{theorem} Let $X$ be an anisotropic quasilinear $p$-hypersurface over $F$, and let $Y$ be a complete variety over $F$ which has no separable points. Suppose that there exists a rational morphism $X \dashrightarrow Y$. Then
\begin{enumerate} \item[$\mathrm{(1)}$] $\mathrm{dim}_{Izh}(X) \leq \mathrm{dim}(Y)$.
\item[$\mathrm{(2)}$] If $\mathrm{dim}_{Izh}(X) = \mathrm{dim}(Y)$, then there exists a rational morphism $Y \dashrightarrow X$. \end{enumerate}
\begin{proof} It follows immediately from Corollary 3.5 that passing to a purely transcendental extension of the base field $F$ changes nothing in the statement. By Lemma 5.5, we may therefore assume that $X$ has a plane section $X'$ of codimension $\mathfrak{i}_1(X) - 1$ with $\mathfrak{i}_1(X') = 1$. Moreover, there exists a rational morphism $X' \dashrightarrow Y$ by Corollary 4.6. Since $\mathrm{dim}_{Izh}(X') = \mathrm{dim}_{Izh}(X)$, we may replace $X$ by $X'$ and assume that $\mathfrak{i}_1(X) = 1$ (or equivalently, $\mathrm{dim}_{Izh}(X) = \mathrm{dim}(X))$. Let $Z$ denote the image of the map $X \dashrightarrow Y$. Since $Y$ has no separable points, the field extension $F(Z)/F$ is not separably generated. It follows from Proposition 3.4 that there exists a rational morphism $Z \dashrightarrow Y'$ for some anisotropic quasilinear $p$-hypersurface $Y'$ over $F$. Now, since the given map $X \dashrightarrow Z$ is dominant, we can consider the composition $f \colon X \dashrightarrow Z \dashrightarrow Y'$. By Corollary 5.11, $X$ is birational to its image in $Y'$ via $f$. It follows that $X$ and $Z$ are birational, and part (1) of the statement follows immediately. Moreover, if we have the equality $\mathrm{dim}(X) = \mathrm{dim}(Y)$, then $X$ and $Y$ are actually birational, whence part (2). \end{proof} \end{theorem}
\begin{remark} Note that the condition that the variety $Y$ has no separable points is equivalent to the condition that $Y$ has no separable \emph{closed} points. Indeed, this follows from the well-known fact that any generically smooth variety has a dense subset of separable closed points. However, the statement of Theorem 5.12 is still weaker than the analogue of Theorem 5.7. More precisely, our result is weaker than the claim that a quasilinear $p$-hypersurface with $\mathfrak{i}_1(X)$ is \emph{strongly p-incompressible} (see Remark 5.8 and \cite{Haution} \S 10). Nonetheless, Theorem 5.12 is sufficient for several interesting applications. \end{remark}
We can now determine which subforms of an anisotropic quasilinear $p$-form $\varphi$ become isotropic over the field $F(\varphi)$. The analogue of the first part of the following statement for quadratic forms over fields of characteristic different from 2 is due to A. Vishik (see \cite{Vishik}, Corollary 3):
\begin{corollary} Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$ of dimension $\geq 2$, and let $\psi$ be a subform of $\varphi$. Then $\psi_{F(\varphi)}$ is isotropic if and only if $\mathrm{codim}_{\varphi}(\psi) \leq \mathfrak{i}_1(\varphi) - 1$. In the case where $\mathrm{codim}_{\varphi}(\psi) = \mathfrak{i}_1(\varphi) - 1$, there is only one rational morphism $X_\varphi \dashrightarrow X_\psi$, and moreover, it is dominant.
\begin{proof} The fact that every subform $\psi$ of codimension $\leq \mathfrak{i}_1(\varphi) - 1$ becomes isotropic over $F(\varphi)$ was proved in Lemma 5.3. The converse follows from the first part of Theorem 5.12. If $\mathrm{codim}_{\varphi}(\psi) = \mathfrak{i}_1(\varphi) - 1$, there is only one rational morphism $X_\varphi \dashrightarrow X_\psi$, because otherwise the subspace of isotropic vectors for $\psi_{F(\varphi)}$ would have dimension $\geq 2$ and we could find a codimension $\mathfrak{i}_1(\varphi)$ subform of $\varphi$ which becomes isotropic over $F(\varphi)$. The map is dominant by part (1) of Theorem 5.12. \end{proof} \end{corollary}
We would like to prove an analogue of Theorem 1.3 for quasilinear $p$-hypersurfaces. To this end, the only obstruction is Question 4.4. We will illustrate this with a proof in the case when $p=2$ or $p=3$ (where we have a positive answer to Question 4.4 by Proposition 4.8). The case where $p=2$ was previously proved by B. Totaro in \cite{Totaro1}.
\begin{theorem} Assume that $p=2$ or $p=3$, and let $X$ and $Y$ be anisotropic quasilinear $p$-hypersurfaces over $F$. Suppose that there exists a rational morphism $X \dashrightarrow Y$. Then
\begin{enumerate} \item[$\mathrm{(1)}$] $\mathrm{dim}_{Izh}(X) \leq \mathrm{dim}_{Izh}(Y)$.
\item[$\mathrm{(2)}$] $\mathrm{dim}_{Izh}(X) = \mathrm{dim}_{Izh}(Y)$ if and only if there is a rational morphism $Y \dashrightarrow X$. \end{enumerate}
\begin{proof} Let $Y' \subset Y$ be a plane section of codimension $\mathfrak{i}_1(Y) - 1$. By Corollary 5.14, we have a dominant rational morphism $Y \dashrightarrow Y'$. Proposition 4.8 now implies that there exists a rational morphism $X \dashrightarrow Y'$. By the first part of Theorem 5.12, we get $\mathrm{dim}_{Izh}(X) \leq \mathrm{dim}(Y') = \mathrm{dim}_{Izh}(Y)$, which proves (1). If there also exists a rational morphism $Y \dashrightarrow X$, then the same argument shows that $\mathrm{dim}_{Izh}(Y) \leq \mathrm{dim}_{Izh}(X)$, and hence $\mathrm{dim}_{Izh}(X) = \mathrm{dim}_{Izh}(Y)$. On the other hand, if we are given the equality $\mathrm{dim}_{Izh}(X) = \mathrm{dim}_{Izh}(Y)$, then there is a rational morphism $Y' \dashrightarrow X$ by the second part of Theorem 5.12. Composing with the dominant rational morphism $Y \dashrightarrow Y'$, we get a rational morphism $Y \dashrightarrow X$, and this proves (2). \end{proof} \end{theorem}
\begin{remark} Note that we do not need a positive answer to Question 4.4 in its entirety to prove the analogue of Theorem 1.3. Indeed, we only need the case where $Z$ is a plane section of codimension $\mathfrak{i}_1(Y) - 1$ in $Y$. For this special case, our question is easily seen to be equivalent to:
\begin{question} Let $Y$ be an anisotropic quasilinear $p$-hypersurface over $F$. Is is true that $\mathfrak{i}_1(Y)$ is minimal among all defect indices attained by $Y$ over extensions of the base field where $Y$ becomes isotropic? \end{question}
\noindent It is a well-established fact that $\mathfrak{i}_1(X)$ satisfies the analogous ``generic property'' when $X$ is a smooth anisotropic quadric. In our case, we have a positive answer to Question 5.17 when $p=2$ or $p=3$ by Proposition 4.8. \end{remark}
\section{Further applications to rational morphisms between quasilinear hypersurfaces}
In this section we use Theorem 5.12 to prove some more specific results concerning rational morphisms between quasilinear $p$-hypersurfaces. In particular, we prove analogues of Hoffmann's Theorem 1.1 (Theorem 6.10) and Izhboldin's Theorem 1.2 (Theorem 6.12).\\
Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$ of dimension $\geq 2$, let $j \in [1,\mathfrak{i}_1(\varphi)]$, and let $\psi \subset \varphi$ be a subform of codimension $\mathfrak{i}_1(\varphi) - j$. Then it is easy to see that $\mathfrak{i}_1(\psi) \geq j$. Indeed, it suffices to show that every codimension $j-1$ subform of $\psi$ becomes isotropic over the field $F(\psi)$ (see Proposition 2.2). But any such subform is isotropic over $F(\varphi)$ by Lemma 5.3, and hence isotropic over $F(\psi)$ by Corollary 4.6. Theorem 5.12 now allows us to prove that equality holds:
\begin{proposition} Let $\varphi$ be an anisotropic quasilinear $p$-form over $F$ of dimension $\geq 2$, let $j \in [1, \mathfrak{i}_1(\varphi)]$, and let $\psi \subset \varphi$ be a subform of codimension $\mathfrak{i}_1(\varphi) - j$. Then $\mathfrak{i}_1(\psi) = j$.
\begin{proof} Suppose that $\mathfrak{i}_1(\psi) > j$. Then there is a subform $\sigma \subset \psi$ of codimension $j$ which becomes isotropic over $F(\psi)$. Let $\tau$ be a subform of $\psi$ of codimension $j-1$ which contains $\sigma$ as a codimension 1 subform. Then there exists a rational morphism $X_\tau \dashrightarrow X_\sigma$ by Corollary 4.6. Since $\tau$ has codimension $\mathfrak{i}_1(\varphi) - 1$ in $\varphi$, there is a dominant rational morphism $X_\varphi \dashrightarrow X_\tau$ by Corollary 5.14. But then taking the composition of these maps gives a rational morphism $X_\varphi \dashrightarrow X_\sigma$, and this is impossible by Corollary 5.14 \end{proof} \end{proposition}
\begin{remark} The analogue of Proposition 6.1 for quadratic forms in characteristic different from 2 was proved by A. Vishik in \cite{Vishik} prior to Karpenko and Merkurjev's Theorem 1.3. \end{remark}
We will now prove analogues of Theorems 1.1 and 1.2 for quasilinear $p$-hypersurfaces. First we need to introduce the class of \emph{quasi-Pfister neighbours}.
\begin{definition} Let $\varphi$ be a quasilinear $p$-form over $F$ of dimension $\geq 2$, and let $n$ be the unique positive integer satisfying $p^{n-1} < \mathrm{dim}(\varphi) \leq p^n$. We say that $\varphi$ is a \emph{quasi-Pfister neighbour} if $\varphi$ is proportional to a subform of a quasi-Pfister form of dimension $p^n$. In this case, the variety $X_\varphi$ will also be called a \emph{quasi-Pfister neighbour}. \end{definition}
\begin{remark} Quasi-Pfister neighbours are analogous to \emph{Pfister neighbours} in the theory of quadratic forms. For example, if $\varphi$ is a neighbour of a quasi-Pfister form $\pi$ over $F$, then for every field extension $L/F$, $\varphi_L$ is isotropic if and only if $\pi_L$ is isotropic. Indeed, if $\mathrm{dim}(\pi) = p^n$ and $\pi_L$ is isotropic, then Lemma 2.6 implies that the subspace of isotropic vectors for $\pi_L$ has dimension at least $p^n - p^{n-1}$, and therefore must intersect the underlying space of $\varphi_L$ non-trivially. \end{remark}
Recall from Example 5.1 that if $\pi$ is an anisotropic quasi-Pfister form of dimension $p^n$, then $\mathfrak{i}_1(\pi) = p^n - p^{n-1}$. Applying Proposition 6.1 to the case of an anisotropic quasi-Pfister neighbour, we therefore get:
\begin{corollary} Let $\varphi$ be an anisotropic quasi-Pfister neighbour over $F$, and let $n$ be the unique positive integer such that $p^{n-1} < \mathrm{dim}(\varphi) \leq p^n$. Then $\mathfrak{i}_1(\varphi) = \mathrm{dim}(\varphi) - p^{n-1}$. \end{corollary}
Now, the following observation is key here:
\begin{proposition} Let $\varphi$ be an anisotropic quasilinear $p$-form of dimension $\geq 2$ over $F$. Then there exists a field extension $\widetilde{F}/F$ such that $\varphi_{\widetilde{F}}$ is an anisotropic quasi-Pfister neighbour.
\begin{proof} Let $\pi = \widehat{\nu}_F(\varphi)$ be the norm form of $\varphi$. Recall that $\varphi$ is proportional to a subform of $\pi$. Let $n$ be the unique positive integer such that $p^{n-1} < \mathrm{dim}(\varphi) \leq p^n$. If $\mathrm{dim}(\pi) = p^n$, then $\varphi$ is already a quasi-Pfister neighbour (of $\pi$). We can therefore assume that $\mathrm{dim}(\pi) \geq p^{n+1}$. In particular, we have $\mathrm{dim}_{Izh}(X_\pi) \geq p^n - 1$ by Example 5.1. Since $\mathrm{dim}(X_\varphi) \leq p^n - 2$, part (1) of Theorem 5.12 implies that there are no rational morphisms $X_\pi \dashrightarrow X_\varphi$. In other words, $\varphi$ remains anisotropic over the field $F(\pi)$. Now, the anisotropic part of $\pi_{F(\pi)}$ is nothing else but the norm form $\pi' = \widehat{\nu}_{F(\pi)}(\varphi)$ of $\varphi$ over $F(\pi)$. We have $\mathrm{dim}(\pi') < \mathrm{dim}(\pi)$, and since $\varphi_{F(\pi)}$ is anisotropic, $\varphi_{F(\pi)}$ is proportional to a subform of $\pi'$. Repeating this procedure as many times as is necessary, we eventually produce an extension $\widetilde{F}/F$ over which $\varphi$ becomes an anisotropic quasi-Pfister neighbour (of the norm form $\widehat{\nu}_{\widetilde{F}}(\varphi)$). \end{proof} \end{proposition}
\begin{remark} We should point out that the analogous statement for nondegenerate quadratic forms is certainly not true in general (see \cite{HoffmannIzhboldin} for a detailed discussion of this problem for quadratic forms over fields of characteristic different from 2). \end{remark}
\begin{corollary} Let $\varphi$ be an anisotropic quasilinear $p$-form of dimension $\geq 2$ over $F$, and let $n$ be the unique positive integer such that $p^{n-1} < \mathrm{dim}(\varphi) \leq p^n$. Then $\mathfrak{i}_1(\varphi) \leq \mathrm{dim}(\varphi) - p^{n-1}$.
\begin{proof} By Proposition 6.6, there is an extension $\widetilde{F}/F$ such that $\varphi_{\widetilde{F}}$ is an anisotropic quasi-Pfister neighbour. By Corollary 6.5, we therefore have $\mathfrak{i}_1(\varphi_{\widetilde{F}}) = \mathrm{dim}(\varphi) - p^{n-1}$. Since we have a natural embedding $F(\varphi) \subset \widetilde{F}(\varphi_{\widetilde{F}})$, we get
\begin{equation*} \mathfrak{i}_1(\varphi) = \mathfrak{i}_0(\varphi_{F(\varphi)}) \leq
\mathfrak{i}_0(\varphi_{\widetilde{F}(\varphi_{\widetilde{F}})}) = \mathfrak{i}_1(\varphi_{\widetilde{F}}) = \mathrm{dim}(\varphi) - p^{n-1}, \end{equation*}
which is what we wanted. \end{proof} \end{corollary}
\begin{example} Let $\varphi$ be an anisotropic form of dimension $p^n + 1$ for some $n \geq 0$. Then $\mathfrak{i}_1(\varphi) = 1$. \end{example}
Now we can prove an analogue of Theorem 1.1 for quasilinear $p$-hypersurfaces:
\begin{theorem} Let $X$ and $Y$ be anisotropic quasilinear $p$-hypersurfaces over $F$. If there exists $n \geq 1$ such that $\mathrm{dim}(Y) \leq p^n - 2 < \mathrm{dim}(X)$, then there are no rational morphisms $X \dashrightarrow Y$.
\begin{proof} By part (1) of Theorem 5.12, it is sufficient to show that $\mathrm{dim}_{Izh}(X) > p^n - 2$, and this follows from Corollary 6.8. \end{proof} \end{theorem}
\begin{remark} The analogue of Corollary 6.8 for quadratic forms over fields of characteristic different from 2 was originally proved by D. Hoffmann as a corollary of Theorem 1.1 (see \cite{Hoffmann1}). Here the roles are reversed. The difference is that we were able to use Theorem 5.12 to prove Corollary 6.6, but, as we have remarked above, the analogue of Corollary 6.6 for nondegenerate quadratic forms is false in general. Corollary 6.8 and Theorem 6.10 were proved in the case $p=2$ by D. Hoffmann and A. Laghribi in \cite{HoffmannLaghribi2} using different methods. \end{remark}
We also get the following analogue of Izhboldin's Theorem 1.2. The case where $p=2$ was proved using different methods by D. Hoffmann and A. Laghribi in \cite{HoffmannLaghribi2}.
\begin{theorem} Let $X$ and $Y$ be anisotropic quasilinear $p$-hypersurfaces over $F$ with $\mathrm{dim}(Y) = p^n - 1 \leq \mathrm{dim}(X)$ for some $n \geq 0$. Suppose that there exists a rational morphism $X \dashrightarrow Y$. Then there exists a rational morphism $Y \dashrightarrow X$. If in addition we have $\mathrm{dim}(X) = p^n -1$, then $X$ and $Y$ are birational.
\begin{proof} Let $X' \subset X$ be a plane section of dimension $p^n - 1$. By Corollary 4.6, there exists a rational morphism $X' \dashrightarrow Y$. By Example 6.9, we have that $\mathfrak{i}_1(X') = 1$. It then follows from Corollary 5.11 that $X'$ is birational to $Y$, whence the result. \end{proof} \end{theorem}
Note that in the case where $\mathrm{dim}(X) = \mathrm{dim}(Y) = p^n -1$ for some $n \geq 0$, we get the stronger assertion (in comparison with Theorem 1.2) that $X$ and $Y$ are birational. For non-quasilinear quadrics, this is still an open problem (see also Conjecture 7.1).
\begin{remark} In a recent article \cite{Haution}, O. Haution has shown that any degree $p$ hypersurface of dimension $p-1$ (over an arbitrary field) which has no closed points of degree prime to $p$ is strongly $p$-incompressible. In particular, an anisotropic quasilinear $p$-hypersurface of dimension $p-1$ is strongly $p$-incompressible. For such varieties, this statement is stronger than the results established above in the present article (see Remark 5.13). The proof uses an extension of K. Zainoulline's degree formula for the Euler characteristic (see \cite{Zainoulline}) to fields of arbitrary characteristic, and also to non-smooth varieties of sufficiently small dimension. The Euler characteristic (of the structure sheaf) does not, however, distinguish between degree $p$ hypersurfaces of dimension $\geq p-1$, so this invariant can only be used to explain the first level of the ``separation'' exhibited by Theorems 6.10 and 6.12. We remark any smooth degree $p$ hypersurface of dimension $p^n -1$ (for any $n\geq 0$) over a field of characteristic $\neq p$ which has no closed points of degree prime to $p$ is strongly $p$-incompressible by the degree formulas of A. Merkurjev (generalising those of Rost; see \cite{Merkurjev}). It is not known at present if the degree formulas used to prove this hold in arbitrary characteristic. \end{remark}
\section{Birational geometry of quasilinear hypersurfaces}
In this section we consider the extension of the results obtained by B. Totaro on the birational geometry of quadrics in \cite{Totaro1} to quasilinear hypersurfaces of higher degree.\\
An old problem of O. Zariski asks whether two stably birational varieties of the same dimension over a field are actually birational. It is well-known that this is false in general, but in the case where both varieties are smooth anisotropic quadrics, it is still an important open problem. Using the fact that a smooth isotropic quadric is a rational variety, one easily shows that two smooth anisotropic quadrics $X$ and $Y$ over a field are stably birational if and only if there exist rational morphisms $X \dashrightarrow Y$ and $Y \dashrightarrow X$. We can therefore ask the following question for arbitrary quadrics:
\begin{conjecture}[Quadratic Zariski Problem] Let $X$ and $Y$ be anisotropic quadrics of the same dimension over a field $k$. Suppose that there exist rational morphisms $X \dashrightarrow Y$ and $Y \dashrightarrow X$. Is it true that $X$ and $Y$ are birational? \end{conjecture}
In a series of papers (\cite{Totaro3}, \cite{Totaro1}, \cite{Totaro2}), B. Totaro has suggested a new approach to this problem by means of a related conjecture concerning rulings on quadrics. We will say that a variety $X$ over a field $k$ is \emph{ruled} if $X$ is birational to $Y \times \mathbb{P}_k^1$ for some variety $Y$ over $k$. Totaro has observed the following consequence of Vishik's result stating that a quadric with first Witt index equal to 1 is incompressible:
\begin{theorem}[\cite{Totaro1}, Corollary 3.2] Let $X$ be an anisotropic quadric over a field $k$. If $\mathfrak{i}_1(X) = 1$, then $X$ is not ruled. \end{theorem}
Moreover he conjectures:
\begin{conjecture}[\cite{Totaro3}, Conjecture 3.1] Let $X$ be an anisotropic quadric over a field $k$. Then $X$ is ruled if and only if $\mathfrak{i}_1(X) > 1$. \end{conjecture}
This is formulated more precisely as follows:
\begin{conjecture}[\cite{Totaro2}, Conjecture 1.1] Let $X$ be an anisotropic quadric over a field $k$. Then $X$ is birational to $X' \times \mathbb{P}_k^{\mathfrak{i}_1(X) - 1}$ for some subquadric $X' \subset X$ of codimension $\mathfrak{i}_1(X) - 1$. \end{conjecture}
Many results are known on all these problems for quadrics of small dimension. In \cite{Totaro1}, Totaro proves Conjecture 7.4 for the entire class of quasilinear quadrics, and then uses it to prove the quasilinear case of Conjecture 7.1. The same approach should work for quasilinear hypersurfaces of higher degree. The analogue of Theorem 7.2 is trivial in this setting:
\begin{proposition} Let $X$ be an anisotropic quasilinear $p$-hypersurface over $F$. If $\mathfrak{i}_1(X) = 1$, then $X$ is not ruled.
\begin{proof} Suppose that $X$ is birational to $Y \times \mathbb{P}_F^1$ for some variety $Y$ over $F$. Then we can construct a rational morphism $X \dashrightarrow X$ as the composition
\begin{equation*} X \dashrightarrow Y \times \mathbb{P}_F^1 \xrightarrow{pr_Y} Y \hookrightarrow Y \times \mathbb{P}_F^1 \dashrightarrow X, \end{equation*}
where the second map is the canonical projection and the third map is the embedding of $Y$ in $Y \times \mathbb{P}_F^1$ at a rational point of $\mathbb{P}_F^1$. Note that the composition is defined because the third map embeds $Y$ as an effective Cartier divisor in $Y \times \mathbb{P}_F^1$ (see Proposition 4.5). But the resulting map is not surjective by construction, and this contradicts the fact that the only rational morphism from $X$ to itself is the identity (see Example 5.2). \end{proof} \end{proposition}
We can now prove an analogue of Conjecture 7.4 for quasilinear hypersurfaces of higher degree. Given the results of \S 5, the proof for the case $p=2$ given by B. Totaro in \cite{Totaro1} carries over verbatim to all primes $p$. We reproduce the argument here for the reader's convenience:
\begin{theorem} Let $X$ be an anisotropic quasilinear $p$-hypersurface over $F$, and let $X' \subset X$ be a plane section of codimension $\mathfrak{i}_1(X) - 1$. Then $X$ is birational to $X' \times \mathbb{P}_F^{\mathfrak{i}_1(X) - 1}$.
\begin{proof} For simplicity of notation, let us put $r = \mathfrak{i}_1(X)$. Let $\varphi$ be an anisotropic form over $F$ which defines the variety $X$, and let $\psi \subset \varphi$ be a subform of codimension $r-1$ which defines its subvariety $X'$. By Corollary 5.14, there exists a dominant rational morphism $\pi \colon X \dashrightarrow X'$ via which we may view $F(\psi)$ as a subfield of $F(\varphi)$. In particular, the defect index of $\varphi$ over $F(\psi)$ is no more than $r$. On the other hand, it is at least this large by a simple application of Corollary 4.6. Hence $\varphi$ has the same defect index over $F(\psi)$ as it does over $F(\varphi)$. Let $v_0,...,v_{r-1}$ be a basis of the subspace of isotropic vectors for the form $\varphi_{F(\psi)}$. The $v_i$ may be regarded as rational morphisms from $X'$ to the affine hypersurface $\lbrace \varphi = 0 \rbrace$. Define a rational morphism $f \colon X' \times \mathbb{P}_F^{r-1} \dashrightarrow X$ over $F$ by the assignment
\begin{equation*} (x',[\lambda_0:...:\lambda_{r-1}]) \mapsto [\lambda_0 v_0(x') + ... + \lambda_{r-1} v_{r-1}(x')]. \end{equation*}
Now, the identity map $X \rightarrow X$ corresponds to some isotropic line in the space $V_\varphi \otimes _F F(\varphi)$. Since $\varphi$ has the same index over $F(\psi)$ as it does over $F(\varphi)$, there are rational functions $f_i \in F(\varphi)$ such that $[f_0 \cdot (v_0 \circ \pi) + ... + f_{r-1} \cdot (v_{r-1} \circ \pi)]$ is the identity map from $X$ to itself. Define a rational morphism $g \colon X \dashrightarrow X' \times \mathbb{P}_F^{r-1}$ by the assignment
\begin{equation*} x \mapsto (\pi(x),[f_0(x),...,f_{r-1}(x)]). \end{equation*}
By construction, the composition $f \circ g$ is the identity on $X$. Therefore $g$ is a birational isomorphism, and the statement is proved. \end{proof} \end{theorem}
Finally, we make some remarks concerning an analogue of the Quadratic Zariski Problem for quasilinear hypersurfaces of higher degree. Unfortunately, the obstruction here is again Question 4.4. As with Theorem 5.15, we illustrate this with a proof for quasilinear quadrics and cubics (for which we know that Question 4.4 has a positive answer).
\begin{theorem} Assume that $p=2$ or $p=3$, and let $X$ and $Y$ be anisotropic quasilinear $p$-hypersurfaces of the same dimension over $F$. Suppose that there exist rational morphisms $X \dashrightarrow Y$ and $Y \dashrightarrow X$. Then $X$ and $Y$ are birational.
\begin{proof} Let $X' \subset X$ and $Y' \subset Y$ be plane sections of codimensions $\mathfrak{i}_1(X) - 1$ and $\mathfrak{i}_1(Y) - 1$ respectively. By Theorem 7.6, $X$ is birational to $X' \times \mathbb{P}_F^{\mathfrak{i}_1(X) - 1}$ and $Y$ is birational to $Y' \times \mathbb{P}_F^{\mathfrak{i}_1(Y) - 1}$. Moreover, we have $\mathfrak{i}_1(X) = \mathfrak{i}_1(Y)$ by Theorem 5.15 (here we are using the statement of Question 4.4). In order to prove the statement, it therefore suffices to show that $X'$ and $Y'$ are birational. Now, we have rational morphisms $X' \dashrightarrow Y$ and $Y \dashrightarrow Y'$ by Corollary 4.6 and Lemma 5.3 respectively. By Proposition 4.8, we therefore have a rational morphism $X' \dashrightarrow Y'$ (again, we are using the statement of Question 4.4). Since $\mathrm{i}_1(X) = 1$ and $\mathrm{dim}(X') = \mathrm{dim}(Y')$, $X'$ and $Y'$ are birational by Corollary 5.11. \end{proof} \end{theorem}
\begin{remark} As with Theorem 5.15, we do not need a positive answer to Question 4.4 in its entirety, but only a positive answer to Question 5.17. \end{remark}
Still, using some of the results we have obtained, we can give partial results towards a positive solution to the Zariski problem for all primes $p$. By Corollary 5.11, the conjecture is true whenever $\mathfrak{i}_1(X) = 1$. In particular, it is true whenever $\mathrm{dim}(X) = \mathrm{dim}(Y) = p^n - 1$ for some $n \geq 0$ by Example 6.9. We can improve this to include dimensions which are ``sufficiently close'' to the form $p^n - 1$. First we need the case of quasi-Pfister neighbours:
\begin{proposition} Let $X$ and $Y$ be quasi-Pfister neighbours of the same dimension over $F$. Suppose that there are rational morphisms $X \dashrightarrow Y$ and $Y \dashrightarrow X$. Then $X$ and $Y$ are birational.
\begin{proof} Let $X' \subset X$ and $Y' \subset Y$ be plane sections of codimensions $\mathfrak{i}_1(X) - 1$ and $\mathfrak{i}_1(Y) - 1$ respectively. From the proof of Theorem 7.7, the only thing left to check is that we have rational morphisms $X' \dashrightarrow Y'$ and $Y' \dashrightarrow X'$. Now, let $\varphi$ and $\psi$ be quasi-Pfister forms over $F$ such that $X$ is a neighbour of $X_{\varphi}$ and $Y$ is a neighbour of $X_{\psi}$. Then there exist rational morphisms $X' \dashrightarrow X_{\psi}$ and $Y' \dashrightarrow X_{\varphi}$ by Corollary 4.6. But by Corollary 6.5, $X'$ and $Y'$ are also neighbours of $X_{\varphi}$ and $X_{\psi}$ respectively. In particular, for any variety $Z$ over $F$, there exists a rational morphism $Z \dashrightarrow X'$ (resp. $Z \dashrightarrow Y'$) if and only if there exists a rational morphism $Z \dashrightarrow X_{\varphi}$ (resp. $Z \dashrightarrow X_{\psi}$; see Remark 6.4). Therefore we have rational morphisms $X' \dashrightarrow Y'$ and $Y' \dashrightarrow X'$, and the proof is complete. \end{proof} \end{proposition}
\begin{remark} For example, any two neighbours of the same quasi-Pfister hypersurface which have the same dimension are birational. Together with Proposition 6.6, Proposition 7.9 shows that the analogue of the quadratic Zariski problem is true up to making an extension of the base field which preserves the anisotropy of the quasilinear hypersurfaces involved. \end{remark}
We conclude with the following result, which settles the Zariski problem in a large number of cases:
\begin{proposition} Let $X$ and $Y$ be anisotropic quasilinear $p$-hypersurfaces of the same dimension $d$ over $F$. Suppose that there exist rational morphisms $X \dashrightarrow Y$ and $Y \dashrightarrow X$, and let $n$ be the unique non-negative integer satisfying $p^n < d+2 \leq p^{n+1}$. If $d \leq p^n + n$, then $X$ and $Y$ are birational.
\begin{proof} Let $\varphi$ and $\psi$ be anisotropic forms defining $X$ and $Y$ respectively. By Proposition 4.10, we have $\mathrm{lndeg}_F(\varphi) = \mathrm{lndeg}_F(\psi)$. Let us denote this integer by $m$. If $m = {n+1}$, then $X$ and $Y$ are quasi-Pfister neighbours, and we are done by Proposition 7.9. We can therefore assume that $m \geq n+2$. Now, let $X' \subset X$ and $Y' \subset Y$ be plane sections of codimensions $\mathfrak{i}_1(X) - 1$ and $\mathfrak{i}_1(Y) - 1$ respectively. As before, we only need to show that there exist rational morphisms $X' \dashrightarrow Y'$ and $Y' \dashrightarrow X'$. By Corollary 4.6 and Lemma 5.3, we have rational morphisms $X' \dashrightarrow Y$ and $Y \dashrightarrow Y'$ (resp. rational morphisms $Y' \dashrightarrow X$ and $X \dashrightarrow X'$), and by Proposition 4.5 it will be sufficient to prove that $Y$ (resp. $X$) is regular at the image of the generic point of $X'$ (resp. $Y'$). We may therefore assume that $X$ and $Y$ are not regular. Note that we have $\mathfrak{i}_1(X') = \mathfrak{i}_1(Y') = 1$ by Proposition 6.1. It therefore follows from Corollary 5.11 that we may also assume that the given rational morphisms $X' \dashrightarrow Y$ and $Y' \dashrightarrow X$ are closed embeddings of subvarieties. Now, the non-regular locus of $X$ (resp. $Y$) has codimension $m \geq n+2$ by Theorem 4.2. On the other hand, the Separation Theorem 6.10 implies that
\begin{equation*} \mathrm{codim}_Y(X') = d - \mathrm{dim}(X') \leq d - (p^n - 1), \end{equation*}
and similarly $\mathrm{codim}_X(Y') \leq d - (p^n - 1)$. Hence if $d \leq p^n + n$, then we have
\begin{equation*} \mathrm{codim}_Y(X'), \mathrm{codim}_X(Y') \leq n+1 < m, \end{equation*}
and so $Y$ (resp. $X$) is regular at the generic point of $X'$ (resp. $Y'$). \end{proof} \end{proposition}
{\bf Acknowledgements.} This work is part of my PhD thesis carried out at the University of Nottingham. I would like to thank Detlev Hoffmann for suggesting several of the problems discussed in this text and for his guidance during the period in which this work was completed. I would also like to thank Olivier Haution and Alexander Vishik for numerous helpful discussions. I am particularly grateful to Alexander Vishik for comments which helped to improve \S 5 of the paper.\\
\bibliographystyle{alphaurl}
|
\section{Introduction}
\label{introduction}
Cataclysmic variables are interacting binaries with a late type secondary star overfilling its Roche lobe and transferring mass from the inner Lagrangian point $L_1$ toward a white dwarf primary. When the white dwarf in not magnetic (or weakly magnetic i.e. an intermediate polar) an accretion disc is formed. Disc matter is transported inwards whereas angular momentum is transported outwards. Accretion is a common phenomenon of a broad family of astrophysical objects. X-ray binaries are the most similar accreting systems to cataclysmic variables. They possess a black hole or a neutron star as central objects. Supermassive black holes are also the central engine which powers active galactic nuclei. Accretion of matter through a disc is the source of brightness variations on a wide range of time scales and energies. The time scale of the variability scales with the mass of the compact object and this is the reason why cataclysmic variables provide excellent laboratories to study accretion physics.
The most characteristic signature of accretion is the flickering activity. These brightness variations, with amplitudes in the range of $0.1 - 1$ magnitude, last from seconds to tens of minutes. There are four basic physical interpretations for the flickering activity that we know of; 1) interaction of the gas stream from the secondary with the outer edge of the disc , 2) turbulent flow in the accretion disc, 3) magnetic dissipation/reconnection in the disc corona and 4) interaction of the inner edge of the disc with the central white dwarf, the so-called boundary layer. Furthermore, a cellular-automaton model was proposed by \citet{yonehara1997} to describe the timing properties of flickering in cataclysmic variables. In this model light fluctuations are produced by occasional flare like events and subsequent avalanche flow in the accretion disc atmospheres. Flares are assumed to be triggered when the mass density exceeds a critical limit. There have been several attempts to localize the source of flickering. For example, enhanced flickering activity during the orbital hump before the eclipse of U\,Gem was observed by \citet{warner1971}. This allowed them to identify the flickering source with the interaction of the gas stream with the outer edge of the disc (hot spot). On the other hand, \citet{bruch1992} studied energies and variability time scales in a large sample of cataclysmic variable and symbiotic systems and concluded that both the disc and the boundary layer are responsible for the flickering. The combination of the inner disc and hot spot is the source of flickering variability in the case of Z\,Cha \citep{bruch1996} and HT\,Cas, V2051\,Oph, UX\,UMa, IP\,Peg \citep{bruch2000}. Eclipse mapping of V2051\,Oph identified the inner disc as the source of high frequency flickering and the stream overflow as the source of low frequency variability \citep{baptista2004}. The inner disc is also responsible for the flickering in the symbiotic system T\,CrB \citep{zamanov1998}. This was studied and confirmed by statistical modeling of turbulent accretion disc flow by \citet{dobrotka2010}. The authors explained the central concentration of flickering by the large number of turbulent eddies required to transfer angular momentum from the inner disc radius outwards. Therefore, the accretion disc appears as the most common source of flickering in cataclysmic variables and related object such as symbiotic systems, specially its inner parts and boundary layer.
Nova likes are a subclass of cataclysmic variables (see \citealt{warner1995} for review) with high mass transfer rates. Their discs are hot and fully ionized most of the time which means that the mass accretion rate is constant through the whole disc (i.e. steady state). This is opposite to dwarf novae where the mass transfer rate is lower, the mass accretion rate is not constant and intersects the critical value of thermal stability somewhere in the disc at some time. This drives a viscous-thermal instability which triggers the dwarf novae phenomenon (see \citealt{kato1998}, \citealt{lasota2001} for review). Dwarf novae discs are not in steady-state and switch between hot (ionized) and cold (non ionized) states with the former presenting smaller disc truncation radii. A small inner disc truncation radius is expected also in the case of nova like systems. The inner disc truncation radius can be produced either by weak magnetic field (i.e. not strong enough to qualify the system as an intermediate polar) or some kind of evaporation (see e.g. \citealt{meyer1994}). The case of ionized steady state discs is well described by the \citet{shakura1973} accretion disc model.
Extensive work has been devoted to localize the source of flickering in cataclysmic variables (see e.g. \citealt{bruch1996}, \citealt{zamanov1998}, \citealt{bruch2000}, \citealt{baptista2004}), but we know little about the physical phenomenon responsible for its variability. The location of flickering in the inner disc and boundary layer incorporates three possible physical processes: turbulence, magnetic reconnection and interaction of the inner disc with the central star. In this paper we are trying to test whether turbulence is the main mechanism responsible for flickering by applying the statistical model of turbulent accretion flow developed in \citet{dobrotka2010}. The model is applied to the flickering activity observed in the nova likes KR\,Aur and UU\,Aqr. In Sec.~\ref{systems} we summarize the observational flickering properties of KR\,Aur and UU\,Aqr. In Sec.~\ref{modelling} we retrace the statistical turbulence model of Dobrotka et al. (2010). The uncertainty of the measured parameters is discussed in Sec.~\ref{errors}. The final results are presented in Sec.~\ref{results}, discussed in Sec.~\ref{discussion} and summarized in Sec.~\ref{summary}.
\section{Observational results}
\label{systems}
\subsection{KR\,Aur}
\label{kraur}
KR\,Aur is a nova like system of VY\,Scl type. Its flickering activity was studied by \citet{kato2002} during a high state. The authors used 17 observations, with a mean coverage of 9000\,s and 50\,s time resolution, to calculate the power density spectra (PDS). The PDS of the flickering is well described by a broken power law, characteristic of red noise. The power law has a slope with index of -1.63 and a cut off frequency at 31.6\,d$^{-1}$ (or 1.5\,d$^{-1}$ in logarithmic units).
Quasi-periodic variability was detected on some nights with a characteristic time scale of $10 - 15$\,min. No superhumps or other coherent signals were observed.
\subsection{UU\,Aqr}
\label{uuaqr}
Photometric light curves of the nova like UU\,Aqr were obtained by \citet{baptista2008} with 10\,s time resolution and a mean duration of 8000\,s per run. The PDS shows a power law with index -1.5 and a break frequency at $\sim 12.96$\,d$^{-1}$ (or 1.11 in logarithmic units). The eclipse mapping technique was applied to analyze the short term variability. Low and high frequency flickering maps are dominated by emission from two asymmetric arcs similar to the case of the dwarf nova IP\,Peg \citep{baptista2002}. The interpretation of the flickering activity in UU\,Aqr is then similar to IP\,Peg, i.e. tidally-induced spiral shock waves in the outer regions of a large accretion disc. Baptista et al. (2002) have suggested that the flickering in UU\,Aqr is caused by turbulence generated by the disc-spiral wave interaction.
\section{Flickering modeling and simulations}
\label{modelling}
\citet{dobrotka2010} developed a statistical method to model the flickering light curves using turbulence. The method is based on angular momentum transport between two adjacent concentric rings in a disc. The angular momentum is transported by discrete turbulent spherical bodies with a dimension scale $x$ described by an exponential distribution function. The dimension scale is proportional to the radial propagation of the turbulent body in the accretion disc. The maximum dimension is the scale height of the disc, while the minimum is limited by zero.
The basic idea is that the sum of the angular momentum difference $\Delta l(r)$ (where $r$ is the distance from the disc center) of all spherical bodies in two adjacent rings must be equal to the global angular momentum difference between the two rings $\Delta L(r)$, i.e. $\Delta l(r) \sim \Delta L(r)$. Since the scale height, density, radial and tangential velocity of the matter in the rings is a function of $r$, the angular momentum differences must be correlated with a certain parameter $k(r)$ which depends on the radial position of every pair of rings, and is given by
\begin{equation}
k(r) = \frac{\Delta L(r)}{\Delta l(r)}.
\label{correlation_coef}
\end{equation}
Therefore, $k(r)$ is a measure of the quantity of events between two adjacent rings at radial distance $r$. This parameter is needed to estimate the number of flares with duration $t \sim x/v_{\rm r}(r)$, where $v_{\rm r}(r)$ is the radial viscous velocity. The number of events with a certain duration is summed up through all pair of rings at distances $r_{\rm i}$
\begin{equation}
\zeta(t) = \sum\limits_{i=1}^{n} k(r_{\rm i}) f(r_{\rm i},t).
\end{equation}
where $f(r_{\rm i},t)$ is the exponential distribution of turbulent eddy sizes with $t \sim x/v_{\rm r}(r)$ at distance $r_{\rm i}$ from the disc center and $n$ is the number of ring pairs. By selecting the number of flares with durations $t$ between a minimal and maximal value and a sampling time step, we get the final histogram used in the synthetic light curve generation (see \citealt{dobrotka2010} for details).
The disc radial profile is based on the \citet{shakura1973} model. Therefore, the disc must be in the hot ionized steady state, characteristic of dwarf novae outbursts (see \citealt{lasota2001} for review). Nova like systems are adequate for such modeling because of their high mass transfer rates which tend to keep them in the hot branch of dwarf novae activity (see \citealt{warner1995} for review).
We used this method to test whether the turbulence scenario is able to fit the observed PDS with realistic values of the disc $\alpha$ parameter and the inner disc truncation radius. The outer disc radius $R_{\rm out}$ is taken as half\footnote{The tidal perturbation of the secondary truncates the outer disc radius at $\sim 0.9 \times R_{\rm L1}$ and hence provides an upper limit to $R_{\rm out}$, but our simulations are not sensitive to this parameter (see \citealt{dobrotka2010})} radius of the primary Roche lobe $R_{\rm L1}$ \citep{paczynski1971};
\begin{equation}
R_{\rm L1} = 0.462~a~\left( \frac{M_{\rm 1}}{M_{\rm 1} + M_{\rm 2}} \right)^{1/3}\,({\rm cm}),
\end{equation}
where $M_{\rm 1}$ and $M_{\rm 2}$ are the primary and secondary mass respectively and $a$ is the distance between the two stars calculated from Kepler's Third law
\begin{equation}
a = 3.53 \times 10^{10} \left( \frac{M_{\rm 1}}{{\rm M_{\rm \odot}}} \right)^{1/3} (1 + q)^{1/3} \left( \frac{P_{\rm orb}}{1\,{\rm h}} \right)^{2/3}\,({\rm cm}),
\end{equation}
$q = M_{\rm 2} / M_{\rm 1}$ is the mass ratio and $P_{\rm orb}$ the orbital period.
The critical mass transfer rate for a disc to be in a hot stage is given by equation (\citealt{hameury1998})
\begin{equation}
\dot{M}_{\rm tr} = 9.5 \times 10^{15} \alpha^{0.01} \left( \frac{M_1}{{\rm M_{\odot}}} \right)^{-0.89} \left( \frac{r}{10^{10}\,{\rm cm}} \right)^{2.68}\,{\rm g}\,{\rm s^{-1}}.
\end{equation}
Using $\alpha$ values from 0.01 to 1.00 we obtain a critical mass transfer rates of $7.8-8.2 \times 10^{16}$\,g\,s$^{-1}$ for both KR\,Aur and UU\,Aqr. By increasing the outer disc radius to $\sim0.9 \times R_{\rm L1}$ the critical mass transfer rate rises to $3.8 - 4.0 \times 10^{17}$\,g\,s$^{-1}$ for KR\,Aur and $3.8 - 3.9 \times 10^{17}$\,g\,s$^{-1}$ for UU\,Aqr. Therefore, we decided to take $8 \times 10^{16}$\,g\,s$^{-1}$ as a lower limit to the mass transfer rate for a steady disc. Simulations were also computed for $1 \times 10^{17}$, $5 \times 10^{17}$ and $1 \times 10^{18}$\,g\,s$^{-1}$. The model input parameters and the PDS values, as reported in the literature, are summarized in Table~\ref{system_parameters}.
\begin{table}
\caption{Measured and calculated parameters. $M_1$ is the primary mass in solar units, $q$ is the mass ratio, $pl$ index is the power law index as measured from the PDS, $f_{\rm ctf}$ the cut off frequency in mHz, $P_{\rm orb}$ the orbital period in hours, $R_{\rm wd}$ the white dwarf radius estimated from \citet{nauenberg1972} in $10^9$\,cm and $R_{\rm out}$ the outer disc radius as half of the primary Roche lobe in units of $10^{10}$\,cm.}
\begin{center}
\begin{tabular}{lcccr}
\hline
\hline
Object & $M_1$ & $q$ & $pl$ index & log$(f_{\rm ctf})$ \\
& (M$_{\rm \odot}$) & & & (d$^{-1}$) \\
\hline
KR\,Aur & 0.59$^{1}$ & 0.6$^{1}$ & -1.63$^{2}$ & 1.50$^{2}$ \\
UU\,Aqr & 0.67$^{3}$ & 0.3$^{3}$ & -1.50$^{4}$ & 1.11$^{4}$ \\
\hline
\hline
Object & $P_{\rm orb}$ & $R_{\rm wd}$ & $R_{\rm out}$ & \\
& (h) & ($10^9$\,cm) & ($10^{10}$\,cm) & \\
\hline
KR\,Aur & 3.907$^{1}$ & 0.87 & 1.87 & \\
UU\,Aqr & 3.900$^{3}$ & 0.80 & 1.95 & \\
\hline
\end{tabular}
\end{center}
References:\\
1 - \citet{shafter1983}, 2 - \citet{kato2002}, 3 - \citet{baptista1994}, 4 - \citet{baptista2008}
\label{system_parameters}
\end{table}
As mentioned above the dimensionless parameter $\alpha$ and the inner disc radius $R_{\rm in}$ are the free parameters of our simulation. We calculated PDS power law indexes ($pl$ hereafter) and cut off frequencies ($f_{\rm ctf}$ hereafter) for the parametric space $\alpha - R_{\rm in}$ with a resolution of $20 \times 20$ using the Lomb-Scargle method \citep{scargle1982}. The $\alpha$ parameter is allowed to vary from 0.01 up to 1.00 which are typical values for cataclysmic variables (see \citealt{warner1995}, \citealt{lasota2001} for review), whereas $R_{\rm in}$ varies from the vicinity of the white dwarf radius (i.e. 1 percent larger) up to $2.0 \times 10^9$\,cm in 20 steps. The white dwarf radius is estimated from the formula given in \citet{nauenberg1972}. For every combination of parameters we calculated a synthetic light curve and its PDS. The synthetic light curves were sampled in identical manner as the observations of KR Aur and UU Aqr reported in \citet{kato2002} and \citet{baptista2008}. These define the frequency limits
and the resolution of the synthetic PDS. We repeated the simulation 1000 times and then calculated the mean PDS. The resulting mean PDS is binned into 20 equally spaced bins and fitted with two power laws. The break frequency $f_{\rm ctf}$ and the red noise slope $pl$ are the main output parameters.
\section{Uncertainty estimate}
\label{errors}
Unfortunately we have no information about the measurement errors of the observed PDS parameters used in this work. The uncertainty must be taken into account because of the limited number of data used in the mean PDS calculation. \citet{kato2002} only used 17 observations while \citet{baptista2008} used 31. This is considerably lower than the 1000 synthetic runs used in our simulations. Furthermore, the length of the observed light curves is variable and some are very short. Our simulated data sets have a constant duration in order to get confident PDS parameters. This can have an impact in the observed PDS properties and must be taken into account in the analysis of the deviation between the observed and simulated data.
In order to estimate this, we performed simulations of 17 synthetic light curves for KR\,Aur with identical duration and sampling as described in Sec.~\ref{kraur} using an ad hoc model with $\dot{M}_{\rm tr} = 1.0 \times 10^{17}$\,g\,s$^{-1}$, $\alpha$ = 0.20 and $R_{\rm in} = 1.0 \times 10^9$\,cm. A mean PDS was calculated from the synthetic light curves and the process was repeated 10000 times. The resulting histograms of the power law index and the logarithm of the cut off frequency are depicted on the upper two panels of Fig.~\ref{simul_hist}, together with their best Gaussian fits. The simulated PDS parameters cluster at the mean values $pl = -1.59$ and ${\rm log}(f_{\rm ctf}) = 1.58$\,d$^{-1}$ with 1-$\sigma$ of 0.06 and 0.12\,d$^{-1}$ respectively. We take this 1-$\sigma$ values as our estimate of the error in the measured PDS parameters. For comparison, we obtain $pl = -1.580 \pm 0.004$ and ${\rm log}(f_{\rm ctf}) = 1.60 \pm 0.01$\,d$^{-1}$ when increasing the number of synthetic light curves to 1000. The corresponding histograms are shown in the inset panels to Fig.~\ref{simul_hist}. A similar analysis was also performed for the case of UU\,Aqr using 31 synthetic light curves per mean PDS as in the observations of \citet{baptista2008}. The duration and resolution of the synthetic light curves is identical to the mean values quoted in Sec.~\ref{uuaqr}. The PDS parameters cluster at the mean values $pl = -1.97$ and ${\rm log}(f_{\rm ctf}) = 1.85$\,d$^{-1}$ with 1-$\sigma$ widths of 0.04 and 0.09\,d$^{-1}$ respectively. For comparison 1000 synthetic light curves per mean PDS yields $pl = -1.961$ and ${\rm log}(f_{\rm ctf}) = 1.836$\,d$^{-1}$ with 1-$\sigma$ widths of 0.004 and 0.007\,d$^{-1}$ respectively. The histograms of the PDS parameters for UU\,Aqr and their best Gaussian fits are depicted in the lower panels of Fig.~\ref{simul_hist}.
\begin{figure*}
\includegraphics[width=60mm,angle=-90]{kraur_pl_scatter_multi.eps}
\includegraphics[width=60mm,angle=-90]{kraur_ctf_scatter_multi.eps}\\
\includegraphics[width=60mm,angle=-90]{uuaqr_pl_scatter_multi.eps}
\includegraphics[width=60mm,angle=-90]{uuaqr_ctf_scatter_multi.eps}
\caption{Histograms of the simulated PDS parameters obtained using an ad hoc model with $\dot{M}_{\rm tr} = 1.0 \times 10^{17}$\,g\,s$^{-1}$, $\alpha$ = 0.20 and $R_{\rm in} = 1.0 \times 10^9$\,cm. Left panels show the histograms of the power law index for KR\,Aur (top) and UU\,Aqr (bottom) whereas right panels present the histograms of the cut off frequency in logarithm units (also KR\,Aur on top and UU\,Aqr at the bottom). The dotted lines show the Gaussian fits to the histograms. Seventeen synthetic runs were produced to generate the PDS of KR\,Aur and 31 runs for UU\,Aqr. The histograms in the inset panels were produced using 1000 synthetic runs. The simulations were repeated 10000 times in each case.}
\label{simul_hist}
\end{figure*}
\section{Results}
\label{results}
Fig.~\ref{results_grid} presents the simulated PDS parameters in the parametric space $\alpha - R_{\rm in}$ using a grey scale. Only the case for $\dot{M}_{\rm tr} = 8 \times 10^{16}$\,g\,s$^{-1}$ is shown for illustration. The isocontours indicate the measured values (middle line) and the $\pm 1$-$\sigma$ errors as estimated in Sec.~\ref{errors}. No contours are visible in the case of UU\,Aqr because they lie outside the plotted $\alpha$ parameter range $0.01 - 1.00$. In this paper we are mainly interested in the $\alpha$ range $0.1 - 0.5$. The lower limit is set by the minimum $\alpha$ parameter required by the disc instability model (DIM) for hot discs while the upper limit is somehow arbitrary. Values higher than 0.5 are theoretically possible but previous works suggest that they are unlikely (see e.g. \citealt{schreiber2003}, \citealt{schreiber2004}, \citealt{king2007}). Simulations of hot discs in cataclysmic variables typically yield $\alpha = 0.1 - 0.2$ whereas observations constrain $\alpha$ to the range $0.1 - 0.4$.
\subsection{KR\,Aur}
\label{results_kraur}
The first clear result of our simulations is that the power law indexes measured from the PDS are in the $\alpha$ interval $0.1 - 1.0$ as required by hot discs of nova like systems in the high state. On the other hand, the cut off frequency points to low $\alpha$ values (see upper right panel in Fig.~\ref{results_grid}). Fig.~\ref{results_grid} also shows that the simulated power law index allows untruncated disc radii while the cut off frequency do suggests a truncated disc (i.e. $R_{\rm in}\ga 2.1 \times 10^{9}$ cm for $\alpha>0.1$). We then decided to search for $R_{\rm in}$ values in the range $\alpha = 0.1 - 0.5$ for which the power law index and cut off frequency overlap within the estimated errors and the results are listed in Table~\ref{results_tab_kraur}. Fig.~\ref{kraur_comp} shows the dependence of the PDS parameters with mass transfer rate in the $R_{\rm in} - \alpha$ plane. For the sake of clarity, Fig.~\ref{results_comparison} also shows the evolution of the mean PDS parameters without the error intervals. The displacement of the isolines towards lower $\alpha$ values with increasing mass transfer rate is clear.
\begin{figure*}
\includegraphics[width=52mm,angle=-90]{kraur_pl_a20xr20_mtr8d16_pds1000.eps}
\includegraphics[width=52mm,angle=-90]{kraur_ctf_a20xr20_mtr8d16_pds1000.eps}\\
\includegraphics[width=52mm,angle=-90]{uuaqr_pl_a20xr20_mtr8d16_pds1000.eps}
\includegraphics[width=52mm,angle=-90]{uuaqr_ctf_a20xr20_mtr8d16_pds1000.eps}
\caption{PDS parameters for KR\,Aur (upper panels) and UU\,Aqr (lower panels) in the parametric space $\alpha - R_{\rm in}$ for $\dot{M}_{\rm tr} = 8 \times 10^{16}$\,g\,s$^{-1}$. Left panels represent the fitted power law and the right panels the logarithm of the cut-off frequency Contour lines indicate the mean value of the parameters and their $\pm 1$-$\sigma$ limits. For UU\,Aqr these lie outside the calculated range $\alpha = 0.01 - 1.00$.}
\label{results_grid}
\end{figure*}
\begin{figure*}
\includegraphics[width=60mm,angle=-90]{kraur_comp_8d16_pds1000.eps}
\includegraphics[width=60mm,angle=-90]{kraur_comp_1d17_pds1000.eps}\\
\includegraphics[width=60mm,angle=-90]{kraur_comp_5d17_pds1000.eps}
\includegraphics[width=60mm,angle=-90]{kraur_comp_1d18_pds1000.eps}
\caption{Evolution of the $\pm 1$-$\sigma$ intervals of the power law (solid line) and the logarithm of the cut off frequency (dotted lines) for KR\,Aur in the parametric space $\alpha - R_{\rm in}$. Four different mass transfer rates are represented in each panel.}
\label{kraur_comp}
\end{figure*}
\begin{table}
\caption{Extreme $\alpha$ and $R_{\rm in}$ values allowed by the overlap of the $\pm1 \sigma$ contours in Fig. 3. Different mass transfer rates are considered.}
\begin{center}
\begin{tabular}{lcccr}
\hline
\hline
$\dot{M}_{\rm tr}$ & min $\alpha$ & min $R_{\rm in}$ & max $\alpha$ & max
$R_{\rm in}$ \\
($10^{17}$\,g\,s$^{-1}$) & & ($10^9$\,cm) & & ($10^9$\,cm) \\
\hline
0.800 & 0.10$^a$ & 0.88$^b$ & 0.40 & 1.56 \\
1.000 & 0.10$^a$ & 0.88$^b$ & 0.40 & 1.60 \\
5.000 & 0.10$^a$ & 0.88$^b$ & 0.33 & 1.64 \\
10.00 & 0.10$^a$ & 0.88$^b$ & 0.25 & 1.67 \\
\hline
\end{tabular}
\end{center}
$^a$ Minimum $\alpha$ parameter allowed by theory.\\
$^b$ Minimum inner disc radius as set by the radius of the white dwarf.\\
\label{results_tab_kraur}
\end{table}
\subsection{UU\,Aqr}
\label{results_uuaqr}
The results of our simulations for the case of UU\,Aqr are totally different than for the previous system. For illustration we show the case of the lower limit mass transfer rate ($8 \times 10^{16}$\,g\,s$^{-1}$) in the lower panels of Fig.~\ref{results_grid}. The observed cut off frequency $\log (f_{\rm ctf}) = 1.11$ is obtained for very low values of the $\alpha$ parameter beyond the upper left corner of the figure (where the simulated value is 1.2, i.e. minimal). Therefore, the whole interval lies outside the calculated $\alpha$ range. Assuming the same trend of $R_{\rm in}$ versus $\alpha$ as in the KR\,Aur case, very large inner disc radii would be required to get a physically realistic $\alpha$ parameter. Therefore, our simulations of the cut off frequency do not yield a satisfactory result. Similarly, the observed power law index can only be reproduced by our simulations for unplausible values $\alpha>1$. And we note that these conclusions do not depend on the assumed mass transfer rate.
The statistical distribution of the simulated cut off frequencies shown in Fig.1 (main figures) are different for KR\,Aur and UU\,Aqr. In the former case the histograms are well fitted with a Gauss function while in the latter case the Gaussian fit only gives a rough approximation. Consequently, our Gaussian-based uncertainties may not be very realistic. However, we note that a minimum error of 5-$\sigma$ is needed to bring the cut off frequency into required (plausible) $\alpha$ interval. On the other hand, the power law histogram clearly follows a Gaussian distribution and even a 10-$\sigma$ error is not able to bring the power law interval into acceptable $\alpha$ values in the UU\,Aqr case. Therefore, we conclude that our results are robust against a reasonable error amplification and hence no acceptable fit can be obtained for UU\,Aqr with our simulations.
\section{Discussion}
\label{discussion}
\subsection{Model degeneracies}
By fitting our turbulence model to the flickering data we find that the best fitted $\alpha$ parameter decreases while $R_{\rm in}$ increases with the selected value of $\dot{M}_{\rm tr}$. How can this be explained? The velocity of matter along the disc radius is given by (see e.g. \citealt{frank1992});
\begin{equation}
v_{\rm r} \sim \alpha^{4/5} \dot{M}_{\rm tr}^{3/10} M_{\rm 1}^{-1/4} r^{-1/4} f^{-14/5},
\label{disk_vr}
\end{equation}
where
\begin{equation}
f = \left[ 1 - \left( \frac{R_{\rm wd}}{r} \right)^{1/2} \right]^{1/4}.
\end{equation}
Equation \ref{disk_vr} shows that the radial viscous velocity $v_{\rm r}$ increases (and hence the flaring time scale decreases) with both $\dot{M}_{\rm tr}$ and $\alpha$. Therefore, the simulations performed with higher $\dot{M}_{\rm tr}$ result in a higher $v_{\rm r}$ and hence a shorter flaring time scale. As a result, the simulated statistics of the flickering changes and so do the PDS parameters, such as the power law slope and break frequency. In order to reproduce the observed power law properties, the drop in flickering time scale or event duration must be compensated by other parameters such as $\alpha$. In other words, the impact of increasing $\dot{M}_{\rm tr}$ can be compensated by decreasing the $\alpha$ parameter in order to keep the observed power law unchanged. This is the reason behind the $\dot{M}_{\rm tr}-\alpha$ correlation. A similar explanation holds for the $\dot{M}_{\rm tr}-R_{\rm in}$ behavior. If we increase $\dot{M}_{\rm tr}$ the characteristic time scale of the flickering events will drop and the modeled flickering statistics will contain a larger number of fast events. In order to eliminate the excess of fast events we need to cut down the inner disc, where short time scale events dominate. Therefore, we can compensate for the increase in $\dot{M}_{\rm tr}$ by increasing $R_{\rm in}$. A correlation between $\alpha$ and $R_{\rm in}$ was already found by \citet{dobrotka2010}. The explanation is again based on the dependence of the radial velocity on $\alpha$. By increasing $\alpha$ the radial velocity also increases which causes a decrease in the event duration. The quantity of short events is then enhanced and the inner disc radius must rise to compensate for it (see Dobrotka et al. 2010 for details).
\subsection{Results compared with DIM predictions}
The study of flickering using the statistical model of \citet{dobrotka2010} yield two different results for the nova like cataclysmic variables KR\,Aur and UU\,Aqr. A successful fit to the observed PDS properties is obtained for KR\,Aur but not UU\,Aqr. In the former case the observed power law index and cut off frequency can be reproduced with a truncated disc and parameter $\alpha \sim 0.1$, as required by the DIM (see \citealt{lasota2001} for review). Accretion disc theory sets a lower limit to $\alpha \simeq 0.1$ while the upper limit depends on the mass transfer rate used in the model. The higher the mass transfer rate, the lower the $\alpha$ parameter. Maximal values of the $\alpha$ parameter $0.25 - 0.40$ are obtained for mass transfer rates in the range $0.08 - 1 \times 10^{18}$\,g\,s$^{-1}$. These $\alpha$ values are in excellent agreement with the ones frequently used or derived in simulations ( $0.1 - 0.2$ , see e.g. \citealt{schreiber2003}, \citealt{schreiber2004}) and with the observed values $0.1 - 0.4$ of fully ionized geometrically thin discs (\citealt{king2007}). On the other hand, the inner disc radius is allowed to vary from nearly the white dwarf radius up to $1.56 - 1.67 \times 10^9$\,cm. The disc in KR\,Aur is found to be untruncated or only marginally truncated. The inner disc radius is found to increase with the mass transfer rate, but it is always considerably smaller than in the case of quiescent dwarf nova (see e.g. \citealt{schreiber2003}), when the disc is not in the steady state, i.e. the mass transfer rate at the inner disc radius is lower than the mass transfer rate from the secondary (i.e. at the outer disc). According to the DIM, the high mass accretion rate during dwarf nova outburst pushes the inner disc radius closer to the primary. The ionized steady state discs of nova like systems posses higher mass transfer rates than dwarf novae discs. Therefore, smaller inner disc radii are expected. The maximum inner disc radius allowed by our simulations ($1.56 - 1.67 \times 10^9$\,cm) are considerably smaller than the value 2.00 $\times 10^9$\,cm used for the dwarf nova SS\,Cyg in quiescence by \citet{schreiber2003}. This is in agreement with the standard accretion disc theory. In summary, our analysis has demonstrated that a simple statistical model of disc turbulence is able to reproduce the observed flickering characteristics of KR\,Aur with a set of disc parameters consistent with DIM theory. The turbulence scenario is then a plausible mechanism and no other processes seem to be required to explain the observed flickering activity.
The case of UU\,Aqr is totally different to that of KR\,Aur. The observed cut off frequency lead to $\alpha$ values lower than 0.01 in the entire range of inner disc radii and for all mass transfer rates. This is unacceptable even for discs in the low state, such as dwarf novae in quiescence. Concerning the power law, our simulations yield to $\alpha$ values larger than the upper limit 1.0 which is not acceptable by basic accretion disc theory. Therefore, no suitable disc turbulence model can succesfully reproduce the observed PDS of UU\,Aqr.
To summarize, we have two similar nova like systems, but the analysis of their flickering properties with our turbulence disc model lead to two completely different results. Is the turbulence scenario wrong for the case of UU\,Aqr? \citet{baptista2008} found spiral structures in the eclipse maps of UU\,Aqr and they claim that the flickering is caused by turbulence generated by the interaction of the disc with the spiral density wave. The model of \citet{dobrotka2010} that we use in this work is based on a disc profile that does not take into account the presence of spiral structures. Furthermore, the model assumes that the largest dimension scale of the turbulent elements is set by the disc scale height. Any other structure like spiral arms can considerably affect the input distribution function of turbulence dimension. We propose that the presence of spiral structures generate turbulence with a totally different distribution function of dimension than in a standard disc and hence our model is not able to describe the situation. Therefore, the turbulence scenario cannot be tested in this system.
Are KR\,Aur and UU\,Aqr so different that one has spiral structures in the disc while the other has not? No eclipse maps of KR\,Aur are available for comparison but
\citet{kato2002} did not find any evidence for superhumps. Superhumps are brightness variations modulated with a time scale close to the orbital period which are caused by disc precession. The physical phenomenon responsible for disc precession and superhump activity is the tidal force of the secondary star. Superhumps are triggered when the disc outer radius exceeds the 3:1 resonance radius. This is satisfied for mass ratios $q < 0.35$ (see e.g. \citealt{whitehurst1991}, \citealt{patterson2005}). Tidal forces do not only generate precessing discs and superhumps but also spiral shocks in the disc (see e.g. \citealt{steeghs2001}). According to the reported binary parameters UU\,Aur, with $q = 0.3$, is likely to be affected by the tidal influence of the secondary but not KR\,Aur with $q = 0.6$. Then different disc behavior may be expected in the two binaries.
\begin{figure}
\includegraphics[width=60mm,angle=-90]{kraur_comp_pds1000.eps}
\caption{Evolution of the power law index (right lines) and cut off frequency (left lines) for KR\,Aur as a function of mass transfer rate marked in units of $10^{17}$\,g\,s$^{-1}$.}
\label{results_comparison}
\end{figure}
\subsection{Distribution of flickering in the disc}
The parameter $k(r)$ provides an estimate of the number of events between two adjacent disc annuli and hence it is a prime indicator of flickering. The lower panel in Fig.~\ref{kraur_disc_prof} shows that $k(r)$ rises very steeply in the inner disc and decreases with increasing $r$. This suggests that the bulk of the events responsible for the flickering activity are located in the inner disc regions which is in agreement with the observations. \citet{dobrotka2010} explained this invoking the maximum dimension scale of the events. The inner disc contains very small turbulent eddies which are able to transfer little angular momentum. Therefore, a large number of eddies is needed. Conversely, the outer disc regions contain very large eddies which are able to transport large amounts of angular momentum. Therefore, a smaller number of events is sufficient.
This is best seen when inspecting the gradient of angular momentum rather than the angular momentum itself. The top panel in Fig.~\ref{kraur_disc_prof} clearly shows that the angular momentum increases toward the outer disc. Flickering is caused by the transport of angular momentum $L(r)$ and hence the difference between adjacent radii is what matters. The gradient of $L(r)$ decreases toward the outer disc, as is illustrated in the middle panel of Fig.~\ref{kraur_disc_prof}. Therefore, the amount of $L$ that needs to be transported decreases toward the outer disc and hence the quantity of turbulent eddies must also decrease (see lower panel in Fig.~\ref{kraur_disc_prof}). In summary, flickering becomes more important at low $r$ values or the innermost part of the disc, as observed (see Sec~\ref{introduction}).
\begin{figure}
\includegraphics[width=59mm,angle=-90]{kraur_disc_prof.eps}
\caption{Disc profile in KR\,Aur. Radial distribution of angular momentum (upper panel), gradient of angular momentum (middle panel) and correlation coefficient $k(r)$ (bottom panel), as computed from equation~\ref{correlation_coef}. The very steep variations at low $r$ values are due to the proximity of the white dwarf. Our simulations are computed at larger radii after the steep variations.}
\label{kraur_disc_prof}
\end{figure}
\section{Summary}
\label{summary}
In this paper we have analyzed the flickering activity of two nova like systems KR\,Aur and UU\,Aqr. We applied a statistical model of flickering simulations developed by \citet{dobrotka2010} based on a simple idea of angular momentum transport in accretion discs. The use of a steady-state disc model requires the disc to be in the hot ionized state. The two selected nova like systems are then adequate targets for this analysis. We successfully fitted the observed power density spectrum of KR\,Aur using an inner disc truncation radius of $0.88 - 1.67 \times 10^9$\,cm and parameter $\alpha \sim 0.10 - 0.40$. Unfortunately, our modeling fails to reproduce the power density spectrum of UU\,Aqr, probably because of the presence of spiral structures detected by \citet{baptista2008}. Such structures are expected in the case of UU\,Aqr because of its low mass ratio $q = 0.3$. On the other hand, KR\,Aur has a larger mass ratio of 0.6 and the disc does not reach the 3:1 resonance radius where tidal forces from the secondary star generate spiral shocks. Therefore, a different disc behavior is naturally expected in the two systems.
Our simulations show that the number of flickering events increases at lower radii and hence concentrate in the inner disc regions. This behavior is explained by the gradient of angular momentum which also increases toward the center of the disc. Furthermore, the duration of the turbulence events depends on the radial viscous velocity which is a function of the mass transfer rate and the $\alpha$ parameter. All these parameters are correlated and, although our modelling provides useful constraints, they cannot be uniquely determined.
\section*{Acknowledgment}
This work is supported in part by the Grand-in-Aid for the global COE programs on "The Next Generation of Physics, spun from Diversity and Emergence" from MEXT. AD was supported also by the Slovak Grant Agency, grant VEGA-1/0520/10. Partly funded by the Spanish MEC under the Consolider-Ingenio 2010 Program grant CSD2006-00070: ''First science with the GTC'' (http://www.iac.es/consolider-ingenio-gtc/).
\vspace{-0.6cm}
\bibliographystyle{mn2e}
|
\section{Introduction}
\label{intro}
Many scientists and engineers use Octave or MATLAB as their preferred programming language. However, dynamic nature of these languages can lead to slower running-time of programs written in these languages compared to programs written in languages which are not as dynamic, like C, C++ and Fortran.
Two dynamic features that contributes to performance issues in major ways are:
\begin{enumerate}
\item Dynamic typing: Types of variables can change during run-time and generally they are not known before run-time.
\item Dynamic resizing: Size of arrays and matrices can change during run-time. Pre-allocation of arrays is not mandatory in Octave/MATLAB and whenever a value is assigned to a location that is not within the range of the array indexes, the array is resized to store the new value.
\end{enumerate}
For example, consider the following function in Octave/MATLAB:
\begin{verbatim}
function z = mmt(x, y)
z=x*y;
end
\end{verbatim}
Type of variables \verb=x= and \verb=y= can be any allowable type. If we are supposed to compile this function statically, we may have to choose the widest possible type for \verb=x= and \verb=y=. This type is an array of complex numbers. The result would be very inefficient code when parameters are of narrower type (e.g. Integers or even arrays of integers). In these cases they are actually wrapped as arrays of complex numbers.
One way to tackle this problem is to use JIT compilers to compile the program or choose the best previously compiled version at run-time. However, this requires run-time overhead and can be really non-trivial. We would like to follow another path in this paper.
We have developed a new language which is similar to Octave/MATLAB in many ways but has a less dynamic nature. In particular, variables must be declared (with their type) before they are defined or used and arrays must be allocated explicitly. We have called this language, tym (Typed MATLAB).\\
Programs written in tym are translated into C++, the generated C++ program uses Octave library and can be called from the Octave \cite{octave} interpreter. To do this, an oct-file should be created by using mkoctfile.\\
Cython \cite{cython} also takes a similar approach to us but it is based on Python which might not be as common as Octave/MATLAB in scientific and engineering communities. OMPC \cite{ompc} is another compiler that translates MATLAB to Python. We have used some of their routines and their grammatical rules in our code.
\section{Overview}
You can see different components that are necessary to translate a program in tym to its equivalent module in C++ in Figure~\ref{tymarch}.\\
Currently, a program should be written as a function in tym. Later, this function is transformed to a C++ module and is compiled to an oct-file. The resulted oct-file can be called from the octave interpreter as a function. This has the advantage that the user can change or write parts of her MATLAB program that is computation-intensive as a function in tym then call it from Octave so she has access to all packages in octave-forge (Neural networks, Image processing, Signal processing, ...) while she programs in tym.\\
Resulted C++ module uses Octave library classes and routines to manipulate matrices and other objects in a way that is compatible with Octave. However, since Octave library does not depend on the Octave interpreter and any C++ program can use it independent of the rest of the Octave, it is also possible to convert a program which is written in tym to an executable that can be run without relying on the octave interpreter.
\begin{figure}
\caption{tym compiler architecture}
\label{tymarch}
\begin{center}
\includegraphics[scale=0.30]{archtym.eps}
\end{center}
\end{figure}
To implement our tym to C++ compiler, we use PLY \cite{ply} which is a Python implementation of compiler construction tools lex and yacc. PLY supports LALR(1) parsing. All grammatical rules should be written in a python module (\verb_tymply.py_ in our compiler). There is a function for every tym programming construct in this module. The grammatical rule which corresponds to a construct is defined in the function's document string. In the body of the function is the action code which is done upon parsing the construct.\\
PLY generates a parser based on the file tymply.py. Since the generated parser is a bottom up LALR(1) parser, we can assume that it does a post-order traversal on Abstract Syntax Tree (AST) and performs some action code upon visiting each AST node. Every AST node is an instance of a tym construct in input text. In our compiler, action codes are responsible for transforming the input program in tym to its equivalent form in C++.\\
AST is an abstract form of parse-tree which is constructed during parsing the input text.
\section{Dynamic Linking and Octave library}
Octave can be dynamically linked with functions which are written in C/C++. Upon linking any of the linked functions can be called from Octave interpreter.\\
Writing Octave compatible functions requires either following the C-Mex interface or using Octave library to manipulate matrices and other objects you would like to pass to Octave. This is necessary to comply to Octave's memory representation of different objects (like matrices). Octave itself uses the Octave library to manipulate matrices and other objects so using it imposes no additional overhead.\\
Using C-Mex interface causes the overhead of converting input parameters of C-Mex routines to its equivalent Octave representation. However, it has the advantage that the functions that uses C-Mex interface can be also linked to MATLAB proprietary software.\\
We have used Octave library so the remainder of this Section is devoted to it and how a C++ function can be linked with Octave.
Consider the example in Figure~\ref{sumsub_cpp}. Using \verb=#include <octave/oct.h>= is necessary so you have access to the required interface to Octava library.\\
\verb=DEFUN_DLD= is a macro that defines the entry point to the function. It has four arguments, name of the function as appear in Octave, list of input parameters to the function, number of output parameters and function's document string.\\
The function should always return an object of type \verb=octave_value_list=. Input parameters are also passed as a \verb=octave_value_list= object (\verb=args=). Input parameters should be extracted from \verb=args= based on their type. Octave passes the input parameters to the function as arrays. However, here, only two integers should be read which are stored in variables \verb=x= and \verb=y= respectively so the first element (indexed by 0) of each input array has been read. Then global variable \verb=error_state= is checked to make sure that input parameters are of expected type. Finally, two values (\verb=a= and \verb=b=) that are supposed to be returned are stored into \verb=retval= which is an \verb=octave_value_list=. Note that operator \verb=()= has been overloaded for \verb=octave_value_list=.\\In order to call this function from Octave, we have to save it as a file that has the same name as the function (\verb=sumsub.cpp= int this example) and run \verb=mkoctfile sumsub.cpp= command. Now we can call this function like any other function from octave interpreter.
\begin{figure}
\caption{A simple function in C++ that can be linked with Octave}
\label{sumsub_cpp}
\begin{footnotesize}
\begin{verbatim}
#include <octave/oct.h>
#include <iostream>
#include <cstdlib>
// File: sumsub.cpp
DEFUN_DLD (sumsub, args, nargout, "do summation and subtraction") {
octave_value_list retval;
if ((args.length()) != 2) {
std::cout<<"invalid number of input params\n";
return retval;
}
int x=args(0).int32_array_value()(0);
int y=args(1).int32_array_value()(0);
if (error_state) {
std::cout<<"invalid type of input parameters\n";
return retval;
}
int a = x + y;
int b = x - y;
retval(0) = a;
retval(1) = b;
return retval;
}
\end{verbatim}
\end{footnotesize}
\end{figure}
The base of all matrices and arrays in octave library is \verb=Array= class. All elements of an array is stored linearly in memory (see \verb=data= in Figure~\ref{array_class}). \verb=len= member keeps the number of elements stored in the array. Dimensions and shape of the array can be achieved by referring to \verb=dimensions= vector. Addresses required for doing lookup or assignment operation is calculated based on this vector. Calculated address is an integer that acts as an index for \verb=data= member. The array can be easily reshaped just by making changes to \verb=dimensions= vector.\\In fact, every \verb=Array= object has a pointer to a \verb=ArrayRep= object (\verb=rep=) that contains members like \verb=data= and \verb=len=. This object also keep the number of \verb=Array= objects that share it in \verb=count=. Every Array increments \verb=count= of its \verb=rep= on construction and decrements it on destruction. In \verb=Array='s destructor it is checked that whether \verb=rep->count= has reached zero and free \verb=rep= when it is so.\\
\verb=Array= is a parametrized type which uses parameter \verb=T= to refer to type of the elements stored in the array. Array of different types can be instantiated by passing different types as \verb=T=.
\begin{figure}
\caption{Definition of Array class in Octave library}
\label{array_class}
\begin{footnotesize}
\begin{verbatim}
template <class T> class Array
{
class ArrayRep {
T *data;
octave_idx_type len;
int count; // reference count
...
}
dim_vector dimensions;
ArrayRep *rep;
...
}
\end{verbatim}
\end{footnotesize}
\end{figure}
\verb=Array= class has methods and other data members which are not shown in Figure~\ref{array_class}. Many of its methods like \verb=resize= and \verb=reshape= should be familiar to any Octave/MATLAB user. Array lookup and assignment is done by calling either the methods \verb=xelem= or \verb=checkelem=. The former does not perform bound checking and the latter does.\\
In Octave library \verb=idx_vector= type is used to represent single indices, range slices and colons. Array slicing is done through two family of methods:
\begin{itemize}
\item \verb=index= methods: Whenever, a sliced array is used in an expression, one of these methods will be called. For example, \verb=a(1:4, 2:6)= can be implemented as \verb=a.index(idx_vector(0, 4), idx_vector(1, 6))=. Note that indexing in Octave library is zero based and upper bounds in slices are exclusive.
\item \verb=assign= methods: Indexed assignment to arrays is done by using one of these methods. For example, \verb_a(1:3, :)=b_ can be implemented as \verb=a.assign(idx_vector(0, 3), idx_vector::colon, b)=.
\end{itemize}
\section{Types and Symbol Table Management}
We have used a stack of symbol tables to implement nested scoping. These symbol tables are necessary to keep track of every identifier's type in input program. Other information like the line number where a variable is declared and the line number where it is defined are also kept to do various checks like whether a variable that is used, is declared and defined before its use.\\
Type information is used for various purposes including resolving ambiguity in the grammar in certain cases and very limited type inference and type checking.
\section{Programming and experimenting with tym}
Currently, you can write a function in tym, translate it to C++ and call it from Octave.\\
In contrast to MATLAB/Octave, variables must be declared before they are used. As I write this paper, there are only five types for variables in tym, namely \verb_int_, \verb_real_, \verb_intArray_ and \verb_realArray_. Types \verb_real_ and \verb_float_ correspond to \verb_double_ and \verb_float_ types in C++. In Octave library, arrays of \verb_double_ (i.e \verb_Array<double>_) are used to represent arrays of floating point numbers so only \verb_realArray_ is available in tym to avoid any extra conversion and copying. Hopefully, support for complex numbers and arrays will be added in the future.
There are also some directives in tym language that tells the tym compiler whether to generate code to do array bound checking, initialization of variables and similar stuff. Whenever they come in the tym program they enable/disable the desired feature from the line afterward.\\
As of this writing three direcitves are available. \verb= $ 'zero_based_arrays'= tells tym compiler that matrices are zero based. \verb= $ 'no_init_vars'= tells the compiler not to generate code for initializing variables and \verb= $ 'no_check_ranges'= make the compiler generate code that does not do bound checking. All these directives are off by default. That is, when no directive has been presented in input program, the compiler generates code for one-based matrices, initializing variables and bound checking which is more consistent with Octave/MATLAB behavior. However, using any of these directive would have positive effect on performance of the generated C++ program, specially \verb= $ 'no_check_ranges'=.
As an example consider the program in Figure~\ref{mult_tym}. This program is a function in tym that multiplies two arrays of type real. All the mentioned directives are used to make it as efficient as possible.
\begin{figure}
\caption{A function in tym that multiplies its parameters}
\label{mult_tym}
\begin{footnotesize}
\begin{verbatim}
$ 'zero_based_arrays'
$ 'no_init_vars'
$ 'no_check_ranges'
function intArray z = mymult(realArray x, realArray y)
int d1x = rows(x)
int d2x = columns(x)
int d1y = rows(y)
int d2y = columns(y)
if (d2x ~= d1y)
error('incompatible dimensions')
end
createArray(z, d1x, d2y)
int i
int j
int k
for i=0:d1x-1
for j=0:d2y-1
z(i, j) = 0
for k=0:d1y-1
z(i, j) = z(i, j) + x(i, k)*y(k, j)
end
end
end
end
\end{verbatim}
\end{footnotesize}
\end{figure}
The tym compiler would translate the function in Figure~\ref{mult_tym} into the C++ code which is shown in Figure~\ref{mult_cxx}. To call this function from Octave you have to make a dynamically loadable Octave module out of it. To do this you can execute \verb=mkoctfile mymult.cpp= in a terminal. The mentioned C++ program can be generated by invoking \verb=python tymc.py mymult.tm=.
\begin{figure}
\caption{Translated version of mymult}
\label{mult_cxx}
\begin{footnotesize}
\begin{verbatim}
#include <octave/oct.h>
#include <iostream>
#include <cstdlib>
DEFUN_DLD (mymult, args, nargout, "") {
octave_value_list retval;
NDArray x=args(0).array_value();
NDArray y=args(1).array_value();
int d1x = x.rows();
int d2x = x.columns();
int d1y = y.rows();
int d2y = y.columns();
if ((d2x != d1y)) {
std::cout<<"error"<<"incompatible dimensions"<<"\n";return retval;
}
int32NDArray z(dim_vector( d1x, d2y));
int i;
int j;
int k;
for (i = (0); i <= (d1x - 1); i += (1)) {
for (j = (0); j <= (d2y - 1); j += (1)) {
z.xelem(i, j) = 0;
for (k = (0); k <= (d1y - 1); k += (1)) {
z.xelem(i, j) =
z.xelem(i, j) + x.xelem(i, k) * y.xelem(k, j);
}
}
}
retval(0) = z;
return retval;
}
\end{verbatim}
\end{footnotesize}
\end{figure}
As an example for slicing see the program in Figure~\ref{addslice_tym} which would be translated to the C++ code shown in Figure~\ref{addslice_cpp} by the tym compiler.
\begin{figure}
\caption{A function that add two array slices in tym}
\label{addslice_tym}
\begin{footnotesize}
\begin{verbatim}
$ 'no_check_ranges'
function intArray z = addslice(intArray x, intArray y)
if (rows(x) < 3 || columns(x) < 3 || rows(y) < 3 || columns(y) < 3)
error('Matrices should be of size at least 3x3')
end
createArray(z, 3, 3)
z = x(1:2, 1:2) + y(2:3, 2:3)
end
\end{verbatim}
\end{footnotesize}
\end{figure}
\begin{figure}
\end{figure}
\begin{figure}
\caption{Translated version of addslice}
\label{addslice_cpp}
\begin{footnotesize}
\begin{verbatim}
#include <octave/oct.h>
#include <iostream>
#include <cstdlib>
DEFUN_DLD (addslice, args, nargout, "") {
octave_value_list retval;
int32NDArray x=args(0).int32_array_value();
int32NDArray y=args(1).int32_array_value();
if ((x.rows() < 3 || x.columns() < 3 || y.rows() < 3 || y.columns() < 3)) {
error("Matrices should be of size at least 3x3");
return retval;
}
int32NDArray z(dim_vector( 3, 3));
z = ((int32NDArray)x.index(idx_vector(1-1, 2-1+1, 1), idx_vector(1-1, 2-1+1, 1)))
+ ((int32NDArray)y.index(idx_vector(2-1, 3-1+1, 1), idx_vector(2-1, 3-1+1, 1)));
retval(0) = z;
return retval;
}
\end{verbatim}
\end{footnotesize}
\end{figure}
You can also try tymc interactively:\\
Type python and then \verb=import tymply as t= in a terminal.
Now you can translate a tym statement to C++ and see the result instantly.\\
Just type \verb=t.yacc.parse("tym_statement")= where \verb=tym_statement= could be any valid tym statement, and press Enter.
\section{Evaluation}
\begin{comment}
size = 100
mymult
Elapsed time is 0.00457 seconds.
mymultcheck
Elapsed time is 0.0622 seconds.
mymultreal
Elapsed time is 0.002945 seconds.
mmymult
Elapsed time is 19.68 seconds.
size = 300
mymult
Elapsed time is 0.1063 seconds.
mymultcheck
Elapsed time is 1.579 seconds.
mymultreal
Elapsed time is 0.06881 seconds.
mmymult
Elapsed time is 515.8 seconds.
size = 500
mymult
Elapsed time is 0.556 seconds.
mymultcheck
Elapsed time is 7.381 seconds.
mymultreal
Elapsed time is 0.8321 seconds.
\end{comment}
We have done done a limited evaluation using different versions of array multiplication. Results of this evaluation is shown in Table~\ref{times}.\\
Three versions of multiplication have been implemented in tym. \verb=mult-real= is the implementation which is shown in Figure~\ref{mult_tym}. \verb=mult-int= uses \verb=intArray= instead of \verb=realArray=. Version \verb=mult-int-check= is similar to \verb=intArray= except that it does not contain directives for not doing bound checking and variable initialization. Version \verb=mult-octave= is an implementation of array multiplication in Octave/MATLAB. First two arrays of size $100 \times 100$ are filled with random numbers and every version is called to do the multiplication. Then two arrays of size $300 \times 300$ are filled with random numbers and the same process is repeated. Results are shown in Table~\ref{times}. As you can see tym versions are far more efficient than Octave/MATLAB version. This is because of dynamic nature of Octave/MATLAB as explained in Section~\ref{intro}.\\
Among versions which are implemented in tym, \verb=mult-int-check= is the slowest one. The main reason is the bound checking that it does for every array look-up and array assignment.\\
It might be expected that \verb=mult-int= be faster than \verb=mult-real= since operations on integers are faster than operations on floating-points. However, \verb=realArray=s are translated into \verb=NDArray=s which are arrays of doubles (i.e \verb=Array<double>=) but \verb=intArray=s are translated into \verb=int32NDArray=s which are arrays of \verb=octave_int32=. \verb=octave_int32= is a wrapper for integer that is provided by Octave to comply with MATLAB's saturation semantic. Several operators are overloaded for \verb=octave_int32= but to achieve efficiency tymc generates code that does not use these overloaded operators. For example \verb_z(i, j) = z(i, j) + x(i, k)*y(k, j)_ would be translated into
\begin{verbatim}
z(i, j) = z(i, j).value() + x(i, k).value() * y(k, j).value()
\end{verbatim}
in case that \verb_x_ and \verb_y_ and \verb_z_ are \verb_intArray_s. In this way, multiplication and addition are done on plain integers (and not wrapped \verb=octave_int32=s) but this also incurs an additional overhead of calling \verb=value= method of \verb=octave_int32= objects whenever it is used in an expression. That might be the major reason for \verb=mult-int= being slower than \verb=mult-real=.
\begin{table}
\caption{Running times of different versions of multiplication.}
\label{times}
\begin{center}
\begin{tabular} {|c|c|c| }
\hline
\textbf{benchmark} & \textbf{$100 \times 100$} & \textbf{$300 \times 300$} \\ \hline
mult-int & 0.00457 & 0.1063 \\ \hline
mult-int-check & 0.0622 & 1.579 \\ \hline
mult-real & 0.002945 & 0.06881 \\ \hline
mult-octave & 19.68 & 155.3 \\ \hline
\end{tabular}
\end{center}
\end{table}
\section*{Acknowledgement}
We also have to credit another work ompc \cite{ompc} which is a MATLAB to python compiler. We have used many of their grammar rules and routines in our work.
\bibliographystyle{plain}
|
\section{Introduction}
Black holes are by far the most celebrated class of solutions derived
from Einstein's field equations. Being among the first types of solutions
to be found almost a century ago, they have undergone an extensive investigation
over the years. Their existence conditions, forms of solutions and set of
properties have been studied in the context of both traditional General Relativity
and generalized gravitational theories admitting either four or more dimensions.
The most well-known example of such a generalized gravitational theory in four
dimensions is provided by the low-energy heterotic string effective theory
\cite{Gross:1986mw,Metsaev:1987zx}. In this theory, the scalar curvature term
of Einstein's theory is only one part of a more complex action functional where
higher-curvature gravitational terms as well as kinetic and interaction terms
of a variety of additional fields (axions, fermions and gauge fields) make their
appearance. The dilatonic Einstein-Gauss-Bonnet theory is a minimal version
of the aforementioned theory and contains the scalar curvature term $R$, a
scalar field called the dilaton, and a quadratic curvature term, the Gauss-Bonnet
(GB) term, given by $R^2_{GB}=R_{\mu\nu\rho\sigma} R^{\mu\nu\rho\sigma}
- 4 R_{\mu\nu} R^{\mu\nu} + R^2$. The GB term in four dimensions can be
expressed as a total derivative term and, normally, makes no contribution to
the field equations; however, the exponential coupling to the dilaton field, that
emerges in the context of the dilatonic Einstein-Gauss-Bonnet theory, ensures
that the GB term is kept in the theory.
In the framework of the dilatonic Einstein-Gauss-Bonnet theory, new types of
black hole solutions emerged that are endowed with non-trivial dilaton hair
(for an indicative list of works on this topic, see \cite{Boulware, Kanti:1995vq,
Guo, Ohta, Maeda,Kleihaus:2011tg}).
The presence of the GB term in the theory
caused the circumvention of the traditional no-hair theorems as it bypassed
the conditions for their validity. In reality, the
existence of the dilatonic black holes violated only the `letter' of the
no-hair theorems, and not the `essence' of them since the dilaton charge
was of a `secondary' type.
However, it became
evident that in the context of this type of generalized gravitational theories
solutions with a much richer structure and a modified set of properties,
compared to the ones in General Relativity, can emerge.
Another class of gravitational solutions whose properties are strictly set by
the General Theory of Relativity are wormholes. They were first discovered
in 1935 as a feature of Schwarzschild geometry \cite{Einstein:1935tc} and
named the `Einstein-Rosen bridge' as they connect two different universes.
Their importance was realised in full when Wheeler \cite{Wheeler:1957mu,Wheeler:1962}
showed that such a bridge, or wormhole, can connect also two distant regions of
our own universe thus opening the way for fast interstellar travel. However,
it was soon demonstrated that this is not possible for the following reasons:
(i) the Schwarzschild wormhole is hidden inside the event horizon of the
corresponding black hole, therefore, is not static but evolves with time;
as a result, the circumference of its `throat' is not constant but opens and
closes so quickly that not even a light signal can pass through
\cite{Kruskal:1959vx,Fuller:1962zza}, (ii) even if a traveler could somehow
pass the throat, she would be bound to exit the wormhole through the past
horizon of the Schwarzschild geometry; this horizon was shown to be
unstable against small perturbations and that it would change to a proper,
and thus non-traversable, horizon at the mere approaching of the traveler
\cite{Redmount:1985,Eardley:1974zz,Wald:1980nk}.
In \cite{Morris:1988cz} a new class of wormhole solutions was found that
possess no horizon and could in principle be traversable. However, some form
of exotic matter whose energy-momentum tensor had to violate all (null,
weak and strong) energy conditions was necessary in order to keep the throat
of the wormhole open. Several studies have considered a phantom field, i.e.
a scalar field with a reversed sign in front of its kinetic term, as a candidate
for such a form of matter
\cite{Ellis:1973yv,Ellis:1979bh,Bronnikov:1973fh,Kodama:1978dw,ArmendarizPicon:2002km}.
It was demonstrated in \cite{Kanti:1995vq} that the GB term leads to an
effective energy-momentum tensor that also violates the energy conditions.
It is in fact this violation that causes the circumvention of the no-hair
theorem forbidding the existence of
regular black holes with non-trivial scalar hair.
It was noted recently in \cite{Bronnikov:2009az}, too,
that the presence of the GB term in the
context of a scalar-tensor theory has the property to evade the various
no-go theorems of General Relativity.
Indeed, various wormhole solutions
were found in the context of gravitational
theories with higher curvature terms
\cite{Hochberg:1990is,Fukutaka:1989zb,Ghoroku:1992tz,Furey:2004rq}.
In the presence of the Gauss-Bonnet term in particular, wormhole solutions
were found in the context of higher-dimensional gravitational theories
\cite{Bhawal:1992sz,Dotti:2006cp,Gravanis,Giribet1,Dotti:2007az,Richarte,Maeda:2008nz,
Giribet2,Simeone}.
In this work, we will investigate the properties of wormhole solutions that arise
in the context of the four-dimensional dilatonic Einstein-Gauss-Bonnet theory,
first reported in \cite{Kanti:2011jz}. The presence of
the higher-curvature GB term, that follows naturally from the compactification
of the 10-dimensional heterotic superstring theory down to four dimensions,
suffices to support these types of solutions without the need for phantom
scalar fields or other forms of exotic matter.
The outline of our paper is as follows: In section II, we present
the theoretical context of our model and discuss the asymptotic forms
of the sought-for wormhole solutions at the regions of radial infinity
and the regular throat. Based on the latter, we derive the embedding diagram
and study the violation of the energy conditions.
We also present a Smarr relation for the wormhole solutions. In section III, we
present the results of our numerical analysis that reveal
the existence of wormhole solutions in the dilatonic Einstein-Gauss-Bonnet
theory, and discuss their properties.
We demonstrate the stability with respect to radial perturbations
of a subset of these solutions in section IV.
In section V we discuss the junction conditions.
The geodesics in these wormhole spacetimes are presented in section VI.
We calculate the magnitude of the acceleration and tidal forces
that a traveler traversing the wormhole would feel in section VII,
and conclude in section VIII.
\section{Einstein-Gauss-Bonnet-Dilaton Theory}
\subsection{Action}
We consider the following effective action
\cite{Mignemi:1992nt,Kanti:1995vq,Chen:2009rv}
motivated by the low-energy heterotic string theory
\cite{Gross:1986mw,Metsaev:1987zx}
\begin{eqnarray}
S=\frac{1}{16 \pi}\int d^4x \sqrt{-g} \left[R - \frac{1}{2}
\partial_\mu \phi \,\partial^\mu \phi
+ \alpha e^{-\gamma \phi} R^2_{\rm GB} \right],
\label{act}
\end{eqnarray}
where
$\phi$ is the dilaton field
with coupling constant $\gamma$, $\alpha $ is a positive numerical
coefficient given in terms of the Regge slope parameter,
and
$R^2_{\rm GB} = R_{\mu\nu\rho\sigma} R^{\mu\nu\rho\sigma}
- 4 R_{\mu\nu} R^{\mu\nu} + R^2$
is the GB correction.
The dilaton and Einstein equations are given by
\begin{eqnarray}
\nabla^2 \phi & = & \alpha \gamma e^{-\gamma \phi}R^2_{\rm GB}
\label{dileq}\\
G_{\mu\nu} & = &
\frac{1}{2}\left[\nabla_\mu \phi \nabla_\nu \phi
-\frac{1}{2}g_{\mu\nu}\nabla_\lambda \phi \nabla^\lambda\phi
\right]
\nonumber\\
& &
-\alpha e^{-\gamma \phi}
\left[ H_{\mu\nu}
+4\left(\gamma^2\nabla^\rho \phi \nabla^\sigma \phi
-\gamma \nabla^\rho\nabla^\sigma \phi\right) P_{\mu\rho\nu\sigma}
\right]
\label{Einsteq}
\end{eqnarray}
with
\begin{eqnarray}
H_{\mu\nu} & = & 2\left[R R_{\mu\nu} -2 R_{\mu\rho}R^\rho_\nu
-2 R_{\mu\rho\nu\sigma}R^{\rho\sigma}
+R_{\mu\rho\sigma\lambda}R_\nu^{\ \rho\sigma\lambda}
\right]
-\frac{1}{2}g_{\mu\nu}R^2_{\rm GB} \ ,
\\
P_{\mu\nu\rho\sigma} & = &
R_{\mu\nu\rho\sigma}
+2 g_{\mu [ \sigma} R_{\rho ]\nu}
+2 g_{\nu [ \rho} R_{\sigma ]\mu}
+R g_{\mu [ \rho} g_{\sigma ]\nu} \ .
\label{HP}
\end{eqnarray}
\subsection{Ansatz and equations}
Throughout this paper we consider only static, spherically-symmetric
solutions of the above set of equations.
Thus we can write the spacetime line-element
in the form \cite{Kanti:1995vq}
\begin{equation}
ds^2 = g_{\mu\nu}\,dx^\mu dx^\nu= -e^{\Gamma(r)}dt^2+e^{\Lambda(r)}dr^2
+r^2\left(d\theta^2+\sin^2\theta d\varphi^2 \right) \ .
\label{metricS}
\end{equation}
As was demonstrated in \cite{Kanti:1995vq}, the dilatonic-Einstein-Gauss-Bonnet
(EGBd) theory admits black hole solutions whose gravitational background has the
line-element of Eq.~(\ref{metricS}). It was also shown that further classes of
solutions emerge in the context of the same theory. One of them, in particular,
possesses no curvature singularity and no proper horizon
(with $g_{tt}$ being regular over the whole radial regime).
However, the $g_{rr}$ metric component
as well as the dilaton field showed some pathological behavior
at a finite radius $r=r_0$,
as seen from the expansion near $r_0$ \cite{Kanti:1995vq}
\begin{eqnarray}
e^{-\Lambda(r)} & = & \lambda_1 (r-r_0) + \cdots \ ,
\\
\Gamma'(r) & = & \frac{\gamma_1}{\sqrt{r-r_0}} + \cdots \ ,
\\
\phi(r) & = & \phi_0 + \phi_1 \sqrt{r-r_0} + \cdots \ .
\end{eqnarray}
The absence of any singular behaviour of the curvature invariants at $r_0$ signifies
that the aforementioned pathological behaviour is merely due to the particular
choice of the coordinate system.
In \cite{Kanti:2011jz} we have argued that this class of asymptotically flat
solutions can be brought to a regular form by employing the coordinate
transformation $r^2 = r_0^2 + l^2$. Then, the metric becomes
\begin{equation}
ds^2 = -e^{2\nu(l)}dt^2+f(l)dl^2
+(l^2+r_0^2)\left(d\theta^2+\sin^2\theta d\varphi^2 \right) \ .
\label{metricL}
\end{equation}
The above form is regular and describes a wormhole solution, where $r_0$ is the
radius of the throat. Indeed, in terms of the new coordinate $l$, the expansion at
$l=0$ assumes the form
\begin{eqnarray}
f(l) & = & f_0 + f_1 l + \cdots \ ,
\label{expansion-throat1}
\\
e^{2\nu(l)} & = & e^{2\nu_0}(1 +\nu_1 l) + \cdots\ ,
\label{expansion-throat2}
\\
\phi(l) & = & \phi_0 + \phi_1 l + \cdots \ ,
\label{expansion-throat3}
\end{eqnarray}
where $f_i$, $\nu_i$ and $\phi_i$ are constant coefficients, and shows no pathology.
Both metric functions and the dilaton field remain finite
in this asymptotic regime.
Thus these solutions possess no horizon.
In addition, all curvature invariant quantities -- including the GB term
-- turn out to be finite at $l=0$, a result that demonstrates the absence
of any singularity.
Substitution of the metric Eq.~(\ref{metricL})
into the dilaton equation (\ref{dileq})
and Einstein equations (\ref{Einsteq})
yields a coupled system of ordinary differential equations (ODE's)
for the metric functions and the dilaton field
\begin{eqnarray}
f' + \frac{f (r^2 f + l^2 - 2 r^2)}{l r^2}
& = &
\frac{r^2 f \phi'^2}{4 l}
+2\alpha \gamma\frac{e^{-\gamma\phi}}{l r^2} \left\{
2 (r^2 f - l^2)(\gamma\phi'^2-\phi'')
+ \phi'\left[\frac{f'}{f}\,(r^2 f-3 l^2) +\frac{4 l r_0^2}{r^2}\right]\right\}, \hspace*{0.3cm}\,\,\,
\label{ode1} \\[1mm] && \nonumber\\
\nu' - \frac{r^2 f -l^2}{2 l r^2}
& = &
\frac{\phi'^2 r^2}{8 l}
+2\alpha \gamma\frac{e^{-\gamma\phi}}{l r^2 f}
\,\nu'\phi' (r^2 f- 3 l^2)\,,\label{ode2} \\[1mm] && \nonumber
\\
\nu''+\nu'^2 + \frac{\nu' (2 l f - r^2 f')}{2 r^2 f}
& + &
\frac{ 2 r_0^2 f - l r^2 f'}{2 r^4 f} \label{ode3} \\
& = &
-\frac{\phi'^2}{4}
+2\alpha \gamma\frac{e^{-\gamma\phi}}{r^2 f} \left\{
2 l \left[\nu' (\gamma\phi'^2-\phi'') -\phi' (\nu'^2+\nu'')\right]
+ \nu' \phi' \left(\frac{3lf'}{f}-\frac{2r_0^2}{r^2}\right)\right\},
\nonumber\\[1mm]
\phi'' + \nu'\phi' + \frac{\phi' (4 l f - r^2 f')}{2 r^2 f}
& = &
4\alpha \gamma\frac{e^{-\gamma\phi}}{r^4 f} \left\{
-2 (r^2 f - l^2)(\nu'^2+\nu'')
+\nu'\left[\frac{f'}{f}\,(r^2 f-3 l^2) +\frac{4 l r_0^2}{r^2}\right]\right\}.
\label{ode4}
\end{eqnarray}
Here $r^2=l^2+r_0^2$ and the prime denotes the derivative with respect to $l$.
Equations (\ref{ode1}), (\ref{ode2}) and (\ref{ode3}) follow from the
$tt$, $ll$ and $\theta\theta$ components
of the Einstein equations, respectively, whereas the last equation (\ref{ode4}) follows
from the dilaton equation.
For the numerical computation we `diagonalize'
Eqs.~(\ref{ode1}), (\ref{ode2}) and (\ref{ode4}) with respect to
$f'$, $\nu'$ and $\phi''$.
The remaining equation, Eq.~(\ref{ode3}),
involving also second derivatives of $\nu$ and $\phi$,
is satisfied if the other three equations are fulfilled.
Thus a system of ODE's must be solved
that consists of two first-order equations
for the metric functions $f$ and $\nu$
and a second-order equation for the dilaton field.
\subsection{Expansions}
The expansion near the throat,
Eqs.~(\ref{expansion-throat1})-(\ref{expansion-throat3}),
once substituted into the
set of equations leads to a number of recursive constraints that determine the
higher-order coefficients in terms of the lower ones.
Looking at the lowest order,
we observe that $f_0$, $\nu_0$ and $\phi_0$ are free parameters.
Also,
the set of parameters of the theory includes the radius of the throat $r_0$ and
the value of $\alpha$.
The value of the constant $\gamma$ will be later fixed to 1 for simplicity.
However, not all of the above parameters are actually independent. To start with,
we observe that the field equations remain invariant under the
simultaneous changes
\begin{equation}
\label{sym1}
\phi \rightarrow \phi +\phi_*
\ , \hspace{1cm}
(r,l) \rightarrow (r,l)\ e^{-\gamma \phi_*/2} \,.
\end{equation}
In addition, the following transformation
\begin{equation}
\label{sym2}
\alpha \rightarrow k \alpha
\ , \hspace{1cm} \phi \rightarrow \phi + \frac{\ln k}{\gamma} \,.
\end{equation}
is also a symmetry of the equations. In the light of the above, we conclude
that only one parameter out of the set $(\alpha, r_0, \phi_0)$ is independent.
We may therefore choose to have a zero asymptotic value of the dilaton field at
infinity which entails fixing the value of the dilaton field at the throat $\phi_0$.
The remaining two parameters can be combined to give a dimensionless
parameter $\alpha/r_0^2$ that will be used throughout our analysis. Finally,
among the two parameters associated with the metric functions,
($\nu_0$, $f_0$), again only the latter is independent -- it is only the derivatives
of the metric function $\nu$ that appear in the field equations of motion,
therefore, we may use this freedom to fix the value of $\nu_0$ in order to
ensure asymptotic flatness at radial infinity.
Thus, our class of wormholes is a two-parameter family of solutions
that, in the context of our analysis, have been chosen to be
$f_0$ and $\alpha/r_0^2$.
When the expansion of the metric functions and dilaton field near the throat
(\ref{expansion-throat1})-(\ref{expansion-throat3}) are substituted into the
field equations, we obtain a set of constraints on the higher-order coefficients.
The constraint on the value of the first derivative of the dilaton field at the
throat is particularly interesting and takes the form
\begin{equation}
\phi_1^2 = \frac{f_0(f_0-1)}{2\alpha\gamma^2 e^{-\gamma \phi_0}
\left[f_0-2(f_0-1)\frac{\alpha}{r_0^2} e^{-\gamma \phi_0}\right]}\ .
\label{constraint-phi1}
\end{equation}
As the left-hand-side of the above equation is positive-definite, the same
must hold for the right-hand-side. We may easily see that the expression
inside the square brackets in the denominator remains positive and has
no roots if
\begin{equation}
\frac{\alpha}{r_0^2} < \frac{1}{2} e^{\gamma \phi_0} .
\label{constraint_1}
\end{equation}
This inequality is automatically satisfied for the set of wormhole solutions presented
in the next section, and gives a lower limit on the size of the throat of
the wormhole $r_0$. Then, the positivity of the right-hand-side of
Eq. (\ref{constraint-phi1}) demands that $f_0 \geq 1$. The solutions satisfying $f_0=1$
comprise a boundary in the phase space of the wormhole solutions and their
physical significance will be discussed in the next section.
As already discussed above, at the asymptotic regime of radial infinity,
i.e. as $l \to \infty$, we demand asymptotic flatness for the two metric
functions and a vanishing value of the dilaton field. Then, the asymptotic
expansion at infinity takes the form
\begin{eqnarray}
\nu & \rightarrow & -\frac{M}{l} + \cdots \ ,
\nonumber \\
f & \rightarrow & 1+ \frac{2M}{l} + \cdots \ ,
\\
\phi & \rightarrow & -\frac{D}{l} + \cdots \ .
\nonumber
\end{eqnarray}
In the above, $M$ and $D$ are identified with the mass and dilaton charge of the
wormhole, respectively. We remind the reader that in the case of the black hole
solutions \cite{Kanti:1995vq}, the parameters $M$ and $D$ were related, and that
rendered the dilatonic hair as ``secondary''. However, in the case of the wormhole
solutions these parameters are not related; this result, together with the fact that
the number of independent parameters near the throat is also two, confirms the
classification of this group of solutions as a two-parameter class of solutions.
\subsection{Wormhole geometry}
A general property of a wormhole is the existence of a throat,
i.~e.~a surface of minimal area (or minimal radius for spherically symmetric
spacetimes). Indeed, this property is implied by the form
of the line element (\ref{metricL}) above, with $f(0)$ and $\nu(0)$
finite.
To cast this condition in a coordinate independent way, we define the
proper distance from the throat in the following way
\begin{equation}
\xi = \int_0^l \sqrt{g_{_{ll}}} dl' = \int_0^l \sqrt{f(l')} dl' \ .
\end{equation}
If we impose the condition for a minimal radius $R=\sqrt{l^2+r_0^2}$ at $l=0$, this
translates to
\begin{equation}
\left.\frac{dR}{d\xi}\right|_{l=0} = 0 , \ \ \
\left.\frac{d^2R}{d\xi^2}\right|_{l=0} > 0\,.
\end{equation}
It is easily seen that the first condition is indeed satisfied.
For the second condition we find
\begin{equation}
\left.\frac{d^2R}{d\xi^2}\right|_{l=0} = \frac{1}{r_0 f_0} > 0 \ .
\end{equation}
This gives a coordinate independent meaning to the parameter $f_0$,
\begin{equation}
f_0 = \left[\left. R\frac{d^2R}{d\xi^2}\right|_{l=0}\right]^{-1} \ .
\end{equation}
A particularly useful concept with which we may examine the geometry of a given
manifold is the construction of the corresponding embedding diagram. In the
present case, we consider the isometric embedding of a plane passing through
the wormhole. Due to the spherical symmetry of the solutions, we may simplify
the analysis and choose $\theta=\pi/2$. Then, we set
\begin{equation}
f(l) dl^2 +(l^2+r_0^2) d\vphi^2 = dz^2 + d\eta^2 +\eta^2 d\vphi^2 \ ,
\label{eucemb}
\end{equation}
where $\{z,\eta,\varphi\}$ are a set of cylindrical coordinates
in the three-dimensional Euclidean space $R^3$.
Regarding $z$ and $\eta$ as functions of $l$, we find
\begin{equation}
\eta(l)=\sqrt{l^2+r_0^2} \ , \ \ \ \
\left(\frac{dz}{dl}\right)^2+\left(\frac{d\eta}{dl}\right)^2 = f(l) \ .
\end{equation}
From the last equation it follows that
\begin{equation}
z(l) = \pm \int_0^l \sqrt{f(l')-\frac{l'^2}{l'^2+r_0^2}} \ dl' \ .
\label{z_emb}
\end{equation}
Hence $\{\eta(l),z(l)\}$ is a parametric representation of a slice of the
embedded $\theta=\pi/2$-plane for a fixed value of $\vphi$.
We observe that the curvature radius of the curve $\{\eta(l),z(l)\}$ at $l=0$ is
given by $R_{0} = r_0 f_0$. Thus $f_0$ has a geometric
meaning: $f_0 = R_{0}/r_0$ is the ratio
of the curvature radius to the radius of the throat.
\subsection{Energy conditions}
As was already mentioned in the Introduction, the existence of the wormhole solution
relies on the violation of the null energy condition. The null energy condition holds if
\begin{equation}
T_{\mu\nu} n^\mu n^\nu \geq 0
\end{equation}
for any null vector field $n^\mu$. In the particular case of spherically symmetric
solutions, and if we employ the Einstein equations, this condition can be expressed as
\begin{equation}
-G_0^0+G_l^l \geq 0 \ , \ \ \ {\rm and } \ \ \ -G_0^0+G_\theta^\theta\geq 0 \ .
\label{Nulleng}
\end{equation}
If one or both of the above conditions do not hold in some region of spacetime, then the
null energy condition is violated. By making use of the expansion of the fields near the
throat (\ref{expansion-throat1})-(\ref{expansion-throat3}), we find that, close to $r_0$,
\begin{equation}
\left[-G_0^0+G_l^l\right]_{l=0} = -\frac{2}{f_0 r_0^2} < 0 \ ,
\label{eng_pert_cond}
\end{equation}
provided $e^{2\nu(0)}\neq 0$, i.e.~in the absence of a horizon.
Thus for the wormhole solutions there is always a region close to the throat
where the null energy condition is violated.
On the other hand, from the asymptotic expansion of the solutions at infinity,
we find
\begin{eqnarray}
-G_0^0+G_l^l & \rightarrow & \frac{D^2}{2} \frac{1}{l^4} + {\cal O}(l^{-5}) \ ,
\label{term1_as}\\
-G_0^0+G_\theta^\theta & \rightarrow & 20 \alpha M D \frac{1}{l^6} + {\cal O}(l^{-7}) \ .
\label{term2_as}
\end{eqnarray}
We observe that, for solutions with a positive dilaton charge $D$ (and a positive mass
$M$) the null energy condition is satisfied in the asymptotic region. However, if the
dilaton charge is negative, the null energy condition is violated also in that
asymptotic region. Note that $D \geq 0$ for the black hole solutions, where it
was given by a positive-definite combination of $(\alpha, M, \phi_\infty)$
\cite{Kanti:1995vq}.
However, $D \geq 0$ does not necessarily hold for
all of the wormhole solutions.
\subsection{Smarr relation}
To derive the Smarr-like mass formula
for the wormhole solutions we start with the definition of the Komar mass
\begin{equation}
M = M_{\rm th} +\frac{1}{4\pi} \int_\Sigma{ R_{\mu\nu} \xi^\mu n^\nu} dV
= M_{\rm th} -\frac{1}{4\pi} \int{ R^0_{0}\sqrt{-g}}d^3x \,,
\label{M_adm}
\end{equation}
where $\xi^\mu$ is the timelike Killing vector field,
$\Sigma$ is a spacelike hypersurface,
$n^\nu$ is a normal vector on $\Sigma$
and $dV$ is the natural volume element on $\Sigma$.
Here $M_{\rm th}$ denotes the contribution of the throat,
\begin{equation}
M_{\rm th} = \frac{1}{2} A_{\rm th} \frac{\kappa}{2\pi} \,,
\label{m_th}
\end{equation}
where $A_{\rm th}$ is the area of the throat,
and $\kappa$ is the surface gravity at the throat
\begin{equation}
\kappa^2 = -1/2 (\nabla_\mu \xi_\nu)(\nabla^\mu \xi^\nu)
\,, \quad
\kappa = \frac{e^{\nu_0}}{\sqrt{f_0}}\nu'(0) \,.
\label{kap}
\end{equation}
To obtain the mass formula we express $ R^0_{0}$ in terms of the effective
stress energy tensor
\begin{equation}
R^0_{0} = T^0_{0}-\frac{1}{2} T^\mu_\mu
= T^0_{0}-\frac{1}{2} T^\mu_\mu
+\frac{1}{2\gamma}\left[ \nabla^2 \phi-\gamma\alpha e^{-\gamma\phi}R^2_{\rm GB} \right]\,,
\label{r00_th}
\end{equation}
where we have added a zero in the form of the dilaton equation.
Multiplication by $\sqrt{-g}$ yields
\begin{equation}
R^0_{0}\sqrt{-g} = \left( T^0_{0}-\frac{1}{2} T^\mu_\mu
-\frac{1}{2}\alpha e^{-\gamma\phi}R^2_{\rm GB}\right)\sqrt{-g}
+\frac{1}{2\gamma}\partial_\mu\left(\sqrt{-g} \partial^\mu \phi\right).
\label{r00g_th}
\end{equation}
Substitution of the metric yields a total derivative for the right-hand-side of the above equation.
Thus, after integration, we find
\begin{equation}
-\frac{1}{4\pi} \int{R^0_{0}\sqrt{-g}}d^3x = -\frac{D}{2\gamma}
+4\frac{e^{\nu_0}}{\sqrt{f_0}}\alpha e^{-\gamma\phi_0}\nu'(0)
+\frac{1}{2\gamma} \frac{e^{\nu_0}}{\sqrt{f_0}}
r_0^2 \phi'(0)\left(1+4\frac{\alpha\gamma^2}{r_0^2} e^{-\gamma\phi_0}\right).
\label{int_r00}
\end{equation}
Now substitution into Eq.~(\ref{M_adm}) gives the
Smarr-like formula
\begin{equation}
M = 2 S_{\rm th} \frac{\kappa}{2\pi} -\frac{D}{2\gamma}
+\frac{1}{8\pi\gamma}\int{\sqrt{-g}g^{ll}\frac{d\phi}{dl}
\left(1+2\alpha\gamma^2 e^{-\gamma\phi} \tilde{R}\right)}d^2x ,
\end{equation}
with
\begin{equation}
S_{\rm th} = \frac{1}{4}\int{ \sqrt{h}\left(1+2\alpha e^{-\gamma\phi}
\tilde{R}\right) d^2x }\,.
\end{equation}
Here $h$ is the induced spatial metric on the throat, $\tilde{R}$ is the scalar
curvature of $h$, and the integral is evaluated at $l=0$.
Defining the normal vector $n_0^\mu$ on the surface $l=0$ the mass formula
becomes
\begin{equation}
M = 2 S_{\rm th} \frac{\kappa}{2\pi} -\frac{D}{2\gamma}
+\frac{D_{\rm th}}{2\gamma}\ ,
\label{smarr}
\end{equation}
with
\begin{equation}
D_{\rm th} =\frac{1}{4\pi}\int{\sqrt{h}e^{\nu_0} n_0^\mu \partial_\mu \phi
\left(1+2\alpha\gamma^2 e^{-\gamma\phi} \tilde{R}\right)}d^2x \ .
\end{equation}
According to the above, the Smarr-like mass formula for wormholes
is obtained by replacing the horizon properties
by the corresponding throat properties in
the known EGBd mass formula for black holes
\cite{Kanti:1995vq,Kleihaus:2011tg}. In addition, an extra contribution
appears that may be interpreted as a modified throat dilaton charge,
while the GB modification is of the same type as the GB modification
of the area (or entropy in the case of black holes).
\section{Numerical wormhole solutions}
For the numerical calculations we use the line-element (\ref{metricL}),
since the functions are well behaved at $r_0$. For the representation of
the numerical results we employ the metric (\ref{metricS}) as well.
To solve the ODEs numerically,
we introduce the compactified coordinate $x=l/(1+l)$,
thus mapping the semi-infinite range of $l$ to the finite range of $x$,
i.e.~$0\leq x\leq 1$.
We cover the parameter range $0.002 \leq \alpha \leq 0.128$
and $1.0001 \leq f_0 \leq 20.0$,
keeping $r_0=1$ fixed. Also we set $\gamma = 1$.
\subsection{Metric and dilaton functions}
Let us start the discussion of the solutions by recalling
the boundary conditions for the system of equations, consisting of
two first order and one second order ODEs.
At the throat $l=0$ regularity requires
\begin{equation}
\left\{\phi'^2 - \frac{f(f-1)}{2\alpha\gamma^2 e^{-\gamma \phi}
\left[f-2(f-1)\frac{\alpha}{r_0^2} e^{-\gamma \phi}\right]}
\right\}_{l=0} = 0\,.
\label{BC_phi}
\end{equation}
This boundary condition has to be supplemented by
$f(0) = f_0$ or $\phi(0)=\phi_0$ in order to obtain a specific solution.
We note that
the asymptotic condition $f \to 1$ for $l\to \infty$ is always satisfied.
This can be seen from the asymptotic form of Eq.~(14), $f'+f(f-1)/l =0$,
which has the general solution $f= l/(l+const)$.
This leaves the asymptotic boundary conditions at radial infinity
\begin{equation}
\lim_{l\to \infty}\nu = 0 \,, \hspace{1cm} \lim_{l\to \infty}\phi = 0 \, .
\end{equation}
Wormhole solutions can be found for every value of $\alpha/r_0^2$
below $\alpha/r_0^2 \approx 0.13$.
This upper bound on $\alpha/r_0^2$
translates into a lower bound on the radius of the throat $r_0$
(for a given $\alpha$).
Thus the radius of the throat $r_0$ can be arbitrarily large.
In Figs.~1a,b and c, we show the metric functions $f(l)$ and $\nu(l)$
and the dilaton function $\phi(l)$, respectively, for an indicative set of
wormhole solutions.
We also exhibit the scaled GB term $\alpha R^2_{\rm GB}$ in Fig.~1d.
Close to the throat the functions show a distinct dependence on both
parameters, $f_0$ and $\alpha/r_0^2$.
However, for intermediate values of $l$, the functions
$f(l)$, and likewise the functions $\nu(l)$, corresponding to
the same value of $\alpha/r_0^2$, approach each other and form clusters.
For larger values of $l$, solutions obtained for different values of
$\alpha/r_0^2$ also merge together. The same behaviour is observed for
the GB term whereas for the dilaton function $\phi(l)$ the same tendency
also exists but becomes pronounced at slightly larger values of $l$.
\begin{figure}[t]
\lbfig{f-1}
\begin{center}
\hspace{0.0cm} \hspace{-0.6cm}
\includegraphics[height=.25\textheight, angle =0]{Fig6b-old.eps}
\hspace{0.5cm} \hspace{-0.6cm}
\includegraphics[height=.25\textheight, angle =0]{Fig6c.eps}
\\[1mm]
\hspace{0.3cm} {(a)} \hspace{7.5cm} {(b)} \hspace{2cm} \\[1mm]
\hspace{0.0cm} \hspace{-0.6cm}
\includegraphics[height=.25\textheight, angle =0]{Fig6d.eps}
\hspace{0.5cm} \hspace{-0.6cm}
\includegraphics[height=.25\textheight, angle =0]{Fig6a-old.eps}\\
\hspace{0.3cm} {(c)} \hspace{7.5cm} {(d)} \hspace{2cm}
\end{center}
\caption
(a) The metric function $f(l)$, (b)
the metric function $\nu(l)$, (c)
the dilaton function $\phi(l)$,
and (d) the scaled GB term $\alpha R^2_{\rm GB}$ versus $l$
for wormhole solutions with
the parameter values $f_0=1.1$, $1.5$, $2.0$
and $\alpha/r_0^2=0.02$, $0.05$.}
\end{figure}
In Fig.~\ref{F-2} we visualize the geometry of the wormhole solutions.
In particular, we present as a typical example in Fig.~\ref{F-2}a
the isometric embedding of the solution with $\alpha/r_0^2=0.02$ and $f_0=1.1$
\cite{Kanti:2011jz}.
Here also the curvature radius at the throat, $R_{0} = r_0 f_0$, is shown.
Also for $\alpha/r_0^2=0.02$, the curves $z$ versus $\eta$ near the throat
are shown in Fig.~\ref{F-2}b for several values of $f_0$.
\begin{figure}[t]
\lbfig{F-2}
\begin{center}
\mbox{\includegraphics[height=.27\textheight, angle =0]{nnFig1a.eps}
\includegraphics[height=.27\textheight, angle =0]{Fig1b.eps}}
\end{center}
\hspace{0.5cm} {(a)} \hspace{8.5cm} {(b)} \hspace{2cm} \\[1mm]
\caption
(a) Isometric embedding of the wormhole solution for
$\alpha/r_0^2=0.02$ and $f_0=1.1$ (taken from \cite{Kanti:2011jz}),
(b) $z$ versus $\eta$ for $\alpha/r_0^2=0.02$ and $f_0=1.1$, $2.0$, $20.0$.
}
\end{figure}
We may also define the profile function $b(r)$ via the equation
\begin{equation}
e^{-\Lambda(r)} \equiv 1-\frac{b(r)}{r} \ .
\label{bfun}
\end{equation}
We note that, at the throat, the profile function $b/r_0$ goes to one
and, thus, $g^{rr}$ vanishes.
For small values of $f_0$, the value of the metric function
$\nu$ at the throat keeps decreasing as $f_0 \to 1$.
Therefore, in the limit $f_0 \to 1$, the metric function $g_{tt}$
tends to zero and a horizon emerges, thus recovering the
class of dilatonic black hole solutions \cite{Kanti:1995vq}.
\subsection{Domain of existence}
Let us now explore the domain of existence of these wormhole solutions.
To that end we construct a set of families of wormhole solutions,
where for each family the value of the parameter $\alpha/r_0^2$ is fixed,
while the second parameter $f_0$ varies within a maximal range of
$1 < f_0 < \infty$.
The values of the parameter $\alpha/r_0^2$
cover the range $0 < \alpha/r_0^2 < 0.13$.
For larger values of $\alpha/r_0^2$ no wormhole solutions are found.
In Fig.~\ref{f-3} we present the domain of existence of the wormhole solutions.
Fig.~\ref{f-3}a shows the scaled area of the throat $A/16 \pi M^2$
versus the scaled dilaton charge $D/M$.
The domain of existence is mapped by the families of solutions
obtained for a representative set of fixed values of $\alpha/r_0^2$.
We observe that the domain of existence of the wormhole solutions is bounded by
three curves indicated by asterisks, crosses and dots.
The boundary indicated by asterisks coincides with the EGBd black hole curve
\cite{Kanti:1995vq}
and corresponds to the limit $f_0 \to 1$.
The limit $f_0 \to \infty$ on the other hand is indicated by crosses.
At the third boundary, marked by dots,
solutions are encountered
that are characterized by a curvature singularity.
\begin{figure}[t]
\lbfig{f-3}
\begin{center}
\mbox{\includegraphics[height=.27\textheight, angle =0]{Fig3a.eps}
\includegraphics[height=.27\textheight, angle =0]{Fig3b.eps}}
\hspace{0.5cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\mbox{\includegraphics[height=.27\textheight, angle =0]{Fig4a.eps}
\includegraphics[height=.27\textheight, angle =0]{Fig4b.eps}}
\hspace{0.5cm} {(c)} \hspace{8cm} {(d)} \hspace{2cm} \\[1mm]
\end{center}
\caption
(a) Scaled area $A/16 \pi M^2$ of the throat
versus the scaled dilaton charge $D/M$
for several values of $\alpha/r_0^2$.
The boundaries represent EGBd black holes (asterisks),
limiting solutions with $f_0\to \infty$ (crosses)
and solutions with curvature singularities (dots).
(b) Metric function $f_0$ at the throat
(determining the curvature radius $R_0=r_0 f_0$)
versus the scaled dilaton charge $D/M$;
(c) dilaton function $\phi(0)$ at the throat
versus the scaled dilaton charge $D/M$;
(d) redshift function $\nu(0)$ at the throat
versus $1/f_0$.
}
\end{figure}
Before discussing the three limiting cases in more detail,
let us consider further the boundary conditions at the throat.
We illustrate the metric function $f_0$ at the throat
for the same set of families of wormhole solutions
in Fig.~\ref{f-3}b.
For fixed values of $\alpha/r_0^2$ below the limiting value for EGBd black holes,
$\left. \alpha/r_0^2 \right|_{\rm bh}=0.0507$
\cite{Kanti:1995vq}, $f_0$ covers the full range $1 < f_0 < \infty$.
Beyond this critical value, however, the minimal value of $f_0$ increases
with increasing $\alpha/r_0^2$.
We note that, for a certain intermediate range of $\alpha/r_0^2$,
$f_0$ is not monotonic.
In Fig.~\ref{f-3}c we show the dilaton field $\phi_0$ at the throat
versus the scaled dilaton charge $D/M$ for the same set of solutions.
Clearly, the domain of $\phi_0$ is bounded.
Again we observe that in the limit $f_0 \to 1$ the
black hole values are obtained.
Inspection of the redshift function $\nu_0$ at the throat,
exhibited in Fig.~\ref{f-3}d,
reveals that $-g_{00}(r_0)$
tends to zero as $f_0 \to 1$.
Thus a horizon emerges in this limit.
\subsubsection{Black hole limit}
\begin{figure}[h!]
\lbfig{f-5}
\begin{center}
\mbox{\includegraphics[height=.27\textheight, angle =0]{Fig5a.eps}
\includegraphics[height=.27\textheight, angle =0]{Fig5b.eps}}
\end{center}
\hspace{1.0cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\caption
(a) Profile function $b/r_0$ and (b) dilaton function $\phi$
versus $r/r_0$ for $f_0=1.1$, $1.01$, $1.001$, $1.0001$
and $\alpha/r_0^2=0.02$ together with the limiting black hole functions.}
\end{figure}
To study the black hole limit in more detail
let us consider a sequence of wormhole solutions
approaching the black hole solution.
We demonstrate this limiting behaviour for solutions
with $\alpha=0.02$ in Fig.~\ref{f-5}.
Here, we exhibit a sequence of solutions with values
of $f_0$ tending to one.
Clearly, the profile function $b(r)$ (Fig.~\ref{f-5}a) and
the dilaton function $\phi(r)$ (Fig.~\ref{f-5}b)
tend fast to the limiting black hole solutions.
For $f_0=1.0001$ the wormhole functions
and their black hole counterparts are already very close.
The redshift functions $\nu$ are not distinguishable for this set of solutions
except very close to the throat, where $\nu_0$ diverges in the limit.
\subsubsection{Large $f_0$ limit}
\begin{figure}[h!]
\lbfig{f-6}
\begin{center}
\mbox{\includegraphics[height=.27\textheight, angle =0]{Fig6b.eps}
\includegraphics[height=.27\textheight, angle =0]{Fig6a.eps}}
\hspace*{0.8cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\end{center}
\caption
(a) Metric function $f(l)$ and (b) scaled GB term $\alpha R_{\rm GB}^2$
versus $l$ for
$f_0=10.0$, $15.0$, $20.0$ and $\alpha/r_0^2=0.02$, $0.1$.}
\end{figure}
Next we consider the limit of large $f_0$,
indicated by crosses in Figs.~\ref{f-3}.
As $f_0 \to \infty$,
the mass $M$ and the dilaton charge $D$ assume finite values.
The same holds for the redshift function $\nu_0$
and the dilaton field $\phi_0$ at the throat,
although the derivative of the dilaton field $\phi_0'$
with respect to the coordinate $l$ diverges like $\sqrt{f_0}$,
as seen from the boundary condition Eq.~(\ref{BC_phi}).
As an example we exhibit the function $f(l)$
in Fig.~\ref{f-6}a for two values of $\alpha/r_0^2$
and increasing values of $f_0$.
We observe that for a given value of $\alpha/r_0^2$,
the functions $f(l)$ deviate from each other
only for small values of $l$ close to the throat,
but coincide for larger $l$.
Moreover, the region where the functions coincide increases
with increasing values of $f_0$.
Thus the solutions approach a limiting solution for $f_0 \to \infty$,
which depends on $\alpha/r_0^2$.
This limiting behaviour is even more pronounced
in the scaled GB term $\alpha R_{\rm GB}^2$,
a curvature invariant which is exhibited
in Fig.~\ref{f-6}b for the same set of parameters.
Extrapolating the value of $\alpha R_{\rm GB}^2$ at the throat
for large $f_0$ indicates
that $\alpha R_{\rm GB}^2$ remains finite for $f_0 \to \infty$.
\subsubsection{Singularity limit}
\begin{figure}[h!]
\lbfig{f-7}
\begin{center}
\mbox{(a) \includegraphics[height=.27\textheight, angle =0]{Fig8a.eps}
(b) \includegraphics[height=.27\textheight, angle =0]{Fig7a.eps}}
\mbox{(c) \includegraphics[height=.27\textheight, angle =0]{Fig7b.eps}
(d) \includegraphics[height=.27\textheight, angle =0]{Fig7f.eps}}
\end{center}
\caption
(a) Scaled GB term $\alpha R_{\rm GB}^2$ versus $l$,
(b) metric function $f(l)$, (c) derivative $f'(l)$,
and (d) derivative of the dilaton function $\phi'(l)$
for several values of $f_0$ and $\alpha/r_0^2=0.1$
demonstrating the emergence of a curvature singularity.
}
\end{figure}
Finally we turn to the third boundary curve,
indicated by dots in Figs.~\ref{f-3}.
This boundary emerges when branches of solutions
with fixed $\alpha/r_0^2$ terminate at singular configurations.
We demonstrate that these singular configurations possess a curvature singularity
at a critical value of $l$, $l_{\rm crit}$, in Fig.~\ref{f-7}a.
Here the scaled GB term $\alpha R_{\rm GB}^2$ is shown versus $l$
for $\alpha/r_0^2=0.1$ for a sequence of solutions approaching the
singular configuration.
We observe a sharp peak in the vicinity of $l_{\rm crit}$, which increases in
size and diverges as the singular configuration is approached.
To elucidate this emergence of a curvature singularity at some
finite critical value of $l$,
we show in Fig.~\ref{f-7}b the metric function $f(l)$,
in Fig.~\ref{f-7}c its derivative $f'(l)$,
and in Fig.~\ref{f-7}d the derivative of the dilaton function $\phi'(l)$
for a fixed value of $\alpha/r_0^2=0.1$ and increasing
values of $f_0$ approaching the singular configuration.
We note that all functions are continuous,
but as the singular configuration is approached, their derivatives develop
a discontinuity at some point $l_{\rm crit}$ outside the throat.
In fact the derivative of the function $f$ has a pole at $l_{\rm crit}$.
\subsection{Energy conditions}
We now turn to the energy conditions, Eqs.~(\ref{Nulleng}).
We demonstrate the violation of the null energy condition in Figs.~\ref{f-9}.
Here we show `normalized' quantities in order to emphasize the change of sign.
The normalization factor is
\begin{equation}
N= \sqrt{(T_0^0)^2+(T_l^l)^2+2(T_\theta^\theta)^2} \ .
\end{equation}
This normalization is responsible for the steep rise observed in the figures
for $f_0=1.1$, which occurs when all quantities are close to zero.
The null energy condition is violated
close to the throat for all wormhole solutions,
as seen explicitly from Eq.~(\ref{eng_pert_cond})
and demonstrated for the set of solutions of Fig.~\ref{f-9}a.
However, for large values of $f_0$ and at the same time
small values of $\alpha/r_0^2$
the dilaton charge $D$ can become negative.
In this case the null energy condition is violated also
in the asymptotic region, as seen from Eq.~(\ref{term2_as}).
In fact we observe that
for the larger values of $f_0$ in Fig.~\ref{f-9}b
the combination $-T_0^0+T_\theta^\theta$ is negative everywhere.
Thus in this case the null energy condition is violated in the whole spacetime.
\begin{figure}[t]
\lbfig{f-9}
\begin{center}
\end{center}
\mbox{\includegraphics[height=.27\textheight, angle =0]{Fig9e.eps}
\includegraphics[height=.27\textheight, angle =0]{Fig9f.eps}}
\hspace*{1.0cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\caption
Demonstration of the violation of the null energy condition for
$\alpha/r_0^2=0.02$ and several values of $f_0$.}
\end{figure}
\subsection{Smarr relation}
\begin{figure}[t]
\begin{center}
\mbox{\includegraphics[height=.27\textheight, angle =0]{Smarr_a.eps}
\includegraphics[height=.27\textheight, angle =0]{Smarrg_a.eps}}
\hspace*{1.0cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\caption{Check of the Smarr relation for several sets of wormhole solutions:
(a) mass $M$ versus $1/f_0$ for $\alpha/r_0^2=0.1$, $\gamma=1$,
(b) mass $M$ versus $\gamma$ for $\alpha/r_0^2=0.02$, $f_0=1.4$.}
\lbfig{smarrfig}
\end{center}
\end{figure}
Let us finally turn to the Smarr-like mass relation,
Eq.~(\ref{smarr}).
This mass relation is a perfect check for the
numerical accuracy and thus for the quality of the solutions.
We demonstrate the Smarr relation in Figs.~\ref{smarrfig}
by comparing the mass $M$ obtained from the asymptotic fall-off of the
metric functions with the mass $M_{\rm Smarr}$ obtained
by evaluating the various terms in the Smarr relation Eq.~(\ref{smarr}).
The Smarr-like formula is well satisfied for all sets of solutions.
In Fig.~\ref{smarrfig}a, we keep $\gamma=1$ as in the main body of the paper
and exhibit the masses $M$ and $M_{\rm Smarr}$ versus $1/f_0$
for $\alpha/r_0^2=0.1$.
The relative error in this case is below $3 \cdot 10^{-5}$
(for $\alpha/r_0^2=0.05$ it is below $10^{-5}$,
and for $\alpha/r_0^2=0.02$ it is below $5 \cdot 10^{-6}$).
In Fig.~\ref{smarrfig}b we have fixed $\alpha/r_0^2=0.02$, $f_0=1.4$
and varied $\gamma$ to address the $\gamma$-dependence of the
mass formula for a set of solutions.
Also for these solutions the relative error is small
and, in particular, it is below $4 \cdot 10^{-5}$.
\begin{figure}[h!]
\lbfig{f-smarr}
\begin{center}
\mbox{\includegraphics[height=.27\textheight, angle =0]{Sth.eps}
\includegraphics[height=.27\textheight, angle =0]{fig_kap.eps}}
\hspace{0.5cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\mbox{\includegraphics[height=.27\textheight, angle =0]{fig_sth.eps}
\includegraphics[height=.27\textheight, angle =0]{fig_dth.eps}}
\hspace{0.5cm} {(c)} \hspace{8cm} {(d)} \hspace{2cm} \\[1mm]
\end{center}
\caption
(a) Scaled entropy-analogue throat quantity $S_{\rm th}/4 \pi M^2$
versus the scaled dilaton charge $D/M$
for several values of $\alpha/r_0^2$.
The boundaries represent EGBd black holes (asterisks),
limiting solutions with $f_0\to \infty$ (crosses)
and solutions with curvature singularities (dots).
The shaded areas indicate linear stability (lilac or lower), instability (red or upper),
undecided yet (white) w.r.t.~radial perturbations
(Sect.~IV).
(b) Scaled surface gravity $\kappa r_0$ at the throat
versus the scaled dilaton charge $D/M$;
(c) scaled first term of the mass formula;
(d) scaled third term of the mass formula.
}
\end{figure}
To gain a better understanding of the quantities contributing in the
Smarr formula, we now consider the various throat properties.
First, we exhibit in Fig.~\ref{f-smarr}a the scaled throat quantity
$S_{\rm th}/4 \pi M^2$ versus the scaled dilaton charge $D/M$
for the full domain of existence of the wormhole solutions.
$S_{\rm th}/4 \pi M^2$ resembles in its structure the entropy
of black holes, since it contains the same correction term
to the throat area $A_{\rm th}$ as in the black hole case
to the horizon area.
In the black hole limit, $S_{\rm th}$ assumes the meaning
of the black hole entropy.
The scaled entropy-analogue throat quantity $S_{\rm th}/4 \pi M^2$
is always greater than one. Only in the Schwarzschild black hole limit,
where the dilaton vanishes, it assumes precisely the value one.
A comparison with Fig.~\ref{f-3}a shows, that
the quantity $S_{\rm th}/16 \pi M^2$ is always larger than
the scaled throat area $A/16 \pi M^2$. This was observed before
for black holes \cite{Kanti:1995vq,Kleihaus:2011tg}.
The color coding of Fig.~\ref{f-smarr}a is related to the stability
of the solution, discussed in the next section.
The scaled surface gravity $\kappa r_0$ at the throat
of the wormhole solutions is exhibited in Fig.~\ref{f-smarr}b.
It is bounded, with its lower boundary given by the
black hole values, which remain close to the Schwarzschild value of 1/2.
The product of the quantities $S_{\rm th}$ and $\kappa$
represents the first term of the Smarr formula, up to numerical factors.
We exhibit this first term, divided by the mass, in
Fig.~\ref{f-smarr}c.
Since the first term is always greater than one
(except for the Schwarzschild case),
the remaining pieces of the mass formula must contribute
negatively to cancel this excess.
The second term in the mass formula, $-D/2\gamma$,
is negative, except for the small region with
large $f_0$ and small $\alpha/r_0^2$.
Its contribution to the scaled mass formula can be read off the
horizontal axis of Fig.~\ref{f-smarr}c.
We exhibit the third term of the scaled mass formula
in Fig.~\ref{f-smarr}d.
It contains the scaled throat dilaton charge, $D_{\rm th}$,
which is modified by the GB term,
analogous to the modification of the area by the GB term.
As expected, this contribution is negative for the wormhole solutions.
It vanishes for the black holes, which represent the upper limit
for $D_{\rm th}$.
\section{Stability}
Our starting point for the study
of the stability is the line element for spherically symmetric
solutions
\begin{equation}
ds^2 = -e^{2\tilde{\nu}}dt^2+\tilde{f}dl^2
+(l^2+r_0^2)\left(d\theta^2+\sin^2\theta d\varphi^2 \right) \,.
\label{metricLL}
\end{equation}
Here we consider only the pulsation modes.
Thus we allow the metric and dilaton functions to depend on the radial
coordinate $l$ and the time coordinate $t$,
\begin{equation}
\tilde{\nu} = \tilde{\nu}(l,t) \ , \ \ \ \
\tilde{f} = \tilde{f}(l,t) \ , \ \ \ \
\tilde{\phi} = \tilde{\phi}(l,t) \ .
\label{tdepfun}
\end{equation}
For the study of the stability behaviour of our solutions, we also need
the time-dependent Einstein and dilaton equations.
These were presented in \cite{Kanti:1998} in the context of the stability analysis
of the dilatonic black hole solutions, and thus we refrain from repeating them here
\footnote{In \cite{Kanti:1998}, the line-element (\ref{metricS}) was used for the
derivation of the time-dependent field equations.
The new set of equations with respect to the new coordinate $l$
may be derived under the redefinitions
$\Gamma \equiv 2 \nu$, $\Lambda\equiv \ln[r^2 f/l^2]$
and the change of variable $dl/dr=r/l$.
Also, the following changes should be made
due to the different conventions used: $\phi \rightarrow -\phi$ and
$\alpha' e^{\phi}/(4g^2) \rightarrow \alpha e^{-\gamma \phi}$.}.
Next, we decompose the metric and dilaton functions into the unperturbed
functions and the perturbations
\begin{eqnarray}
\tilde{\nu}(l,t) & = & \nu(l) +\epsilon \delta\nu(l) e^{i\sigma t} \ ,
\label{decomp_nu}\\
\tilde{f}(l,t) & = & f(l) +\epsilon \delta f(l) e^{i\sigma t} \ ,
\label{decomp_f}\\
\tilde{\phi}(l,t) & = & \phi(l) +\epsilon \delta\phi(l) e^{i\sigma t} \ ,
\label{decomp_phi}
\end{eqnarray}
where we assume a harmonic time dependence of the perturbations
and $\epsilon$ is considered as small.
Now we substitute the perturbed functions into the Einstein and dilaton
equations and linearize in $\epsilon$.
This yields a system of linear ODEs for the functions
$\delta\nu(l)$, $\delta f(l)$ and $\delta\phi(l)$, where the
coefficients depend on the unperturbed functions and their derivatives.
We use the unperturbed equations to eliminate $\nu'$, $\nu''$,
$f'$ and $\phi''$.
The $tl$ part of the Einstein equations yields $\delta f$
in terms of $\delta\phi$ and $\delta\phi'$. Thus $\delta f$ and $\delta f'$
can be eliminated from the rest of the ODEs. The dilaton equation and the
$\theta\theta$ part of the Einstein equations can be diagonalized with
respect to $\delta \nu''$ and $\delta \phi''$. Finally
$\delta \nu'$ can be eliminated from the dilaton equation by adding
the $rr$ part of the Einstein equations with a suitable factor.
Thus we end up with a single second order equation for $\delta \phi$,
\begin{equation}
\delta \phi'' + q_1 \delta\phi' + (q_0+q_\sigma\sigma^2) \delta\phi = 0 \ ,
\label{lineq1}
\end{equation}
where the coefficients $q_1$, $q_0$ and $q_\sigma$ depend on the unperturbed solution.
The coefficient $q_\sigma$ tends to one asymptotically
and is bounded at $l=0$.
The coefficients $q_1$ and $q_0$ tend to zero asymptotically, however, they
diverge at $l=0$ as $1/l$. Thus, to obtain solutions which are regular at $l=0$
suitable boundary conditions are needed.
Since the perturbations are assumed to be normalizable, $\delta\phi$
has to vanish asymptotically. In order to obtain a unique solution
we also fix $\delta \phi$ at $l=0$. Thus we are left with three boundary
conditions for a second order ODE, which can be satisfied only for certain
values of the eigenvalue $\sigma^2$.
We can avoid the singularity of $q_1$ and $q_0$ at $l=0$, by employing
the transformation
$\delta \phi = F(l) \psi(l)$, where $F(l)$ satisfies
$F'/F = -q_1(l)/2$. This yields
\begin{equation}
\psi'' + Q_0\psi +\sigma^2 q_\sigma \psi = 0 \ ,
\label{lineq2}
\end{equation}
where the coefficient $Q_0=-q'_1/2-q_1^2/4+q_0$ is bounded at $l=0$.
We note that $F(l)$ diverges like $1/l$ at $l=0$. Thus for acceptable solutions
we have to impose the boundary condition $\psi=0$ at $l=0$. In addition
$\psi$ has to vanish asymptotically to ensure normalized solutions.
However, these boundary conditions do not lead to unique solutions, since
the ODE (\ref{lineq2}) is homogeneous. Therefore, we supplement
Eq.~(\ref{lineq2}) with the auxiliary ODE $N' = \psi^2$ and impose the
boundary conditions $N(0)=0$ and $N \to 1$ asymptotically.
These conditions give exactly the normalization
$\int_0^\infty \psi^2 dl = 1$.
Again, solutions exist only for certain values of the eigenvalue $\sigma^2$.
The ODE (\ref{lineq2}) can be written as a Schr\"odinger equation by
introducing the new coordinate $y$ via $dy/dl = \sqrt{q_\sigma}$,
and eliminating the first derivative $d\psi/dy$ as above.
This yields
\begin{equation}
\frac{d^2\chi}{dy} - V_{\rm eff}\chi +\sigma^2 \chi = 0 \ ,
\label{lineq3}
\end{equation}
where
$$
- V_{\rm eff} =
\frac{1}{2\sqrt{q_\sigma}}\left(\frac{1}{\sqrt{q_\sigma}}\right)''
-\frac{1}{4}\left[\left(\frac{1}{\sqrt{q_\sigma}}\right)'\right]^2
+\frac{Q_0}{q_\sigma} \ ,
$$
and $\chi$ is related to $\psi$. The new potential $V_{\rm eff}$ is
bounded for all $y$.
Note that the derivation of Eq.~(\ref{lineq3}) holds only if the
function $q_\sigma$ is positive for all $l$.
However, if $q_\sigma<0$ on some interval,
one cannot rely on the rule that the smallest eigenvalue
corresponds to the eigenfunction without knots.
Indeed we found in some cases
two different eigenfunctions $\psi_1$, $\psi_2$ without knots which
satisfy the orthogonality condition $\int q_\sigma \psi_1 \psi_2 dl =0$.
In order to determine the change of stability
we consider families of wormhole solutions with fixed $\alpha/r_0^2$.
The ODE (\ref{lineq1}) is then solved
together with the normalization constraint for varying values of $f_0$.
In Fig.~\ref{f-pert}(a) we show the eigenvalue $\sigma^2$ versus $f_0$
for several values of $\alpha/r_0^2$.
We observe that solutions for negative values of $\sigma^2$
exist for large values of $f_0$, corresponding to
an instability of the wormhole solutions.
For these solutions the function $\delta\phi$,
respectively $\psi$, decays exponentially.
Thus these solutions correspond to bound states of the equivalent
Schr\"odinger equation.
As $f_0$ decreases, the eigenvalue increases
and tends to zero for some critical value
$f_0=f_0^{\rm (cr)}$, which depends on $\alpha/r_0^2$.
No solutions were found for $f_0<f_0^{\rm (cr)}$.
Solutions of Eq.~(\ref{lineq1}) with positive
$\sigma^2$ would oscillate in the asymptotic region, and are not normalizable.
\begin{figure}[h!]
\lbfig{f-pert}
\begin{center}
\includegraphics[height=.25\textheight, angle =0]{Fig11b.eps}
\includegraphics[height=.25\textheight, angle =0]{FIG3.eps}
\end{center}
\hspace*{0.5cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\caption
(a) Eigenvalue $\sigma^2$ versus $f_0$ for $\alpha/r_0^2=0.02$.
(b) Domain of existence: scaled area $A/16 \pi M^2$ of the throat
versus the scaled dilaton charge $D/M$
for several values of $\alpha/r_0^2$.
The shaded areas indicate linear stability (lilac or lower), instability (red or upper),
undecided yet (white) w.r.t.~radial perturbations (taken from \cite{Kanti:2011jz}).
}
\end{figure}
The interpretation is that the instability mode vanishes
at some critical value $f_0^{\rm (cr)}$.
Hence we conclude that the wormhole solutions are stable for $f_0<f_0^{\rm (cr)}$.
However, this holds only if $q_\sigma >0$ for all $l$.
Otherwise one cannot be sure that $\sigma^2$ is the smallest possible eigenvalue.
Therefore, eigenfunctions with negative eigenvalues
cannot be excluded even if $\sigma^2=0$.
Consequently, if $q_\sigma$ is not positive for all $l$
the question of stability cannot be decided by considering the
standard equivalent Schr\"odinger eigenvalue problem.
Numerically, we found that $q_\sigma$ is positive for all wormhole solutions
with $\alpha/r_0^2 \leq 0.05$;
but wormhole solutions where $q_\sigma$ is negative on some interval exist
if $\alpha/r_0^2 \geq 0.05$.
In Fig.~\ref{f-pert}(b) we show the stability region in the domain of existence
\cite{Kanti:2011jz}.
Also shown is the line where the eigenvalue $\sigma^2$ changes sign.
The instability region (red or upper) is characterized by negative $\sigma^2$,
whereas the stability region (lilac or lower) is characterized
by the absence of negative eigenvalues and positive $q_\sigma$.
In the remaining (white)
region the question of stability is not yet decided.
\section{Junction Conditions}
Up to this point, we have discussed the behaviour of the metric functions
and of the dilaton field in only half of the wormhole spacetime, i.e.~the part
with $l>0$. Our solutions should naturally be extended to the second asymptotically
flat part of the manifold ($l \rightarrow -\infty$). If this is performed by
demanding that the derivatives of the metric and dilaton functions are continuous,
we observe a singular behaviour corresponding to curvature singularities.
This is demonstrated in Fig.~\ref{nonsymmext-1} for the wormhole solution
with parameters $\alpha/r_0^2=0.02$,
$f_0=1.1$ and $\gamma = 1$.
\begin{figure}[h!]
\lbfig{nonsymmext-1}
\begin{center}
\hspace{0.0cm} \hspace{-0.6cm}
\includegraphics[height=.25\textheight, angle =0]{diffFig2.eps}
\hspace{0.5cm} \hspace{-0.6cm}
\includegraphics[height=.25\textheight, angle =0]{diffFig3.eps}
\end{center}
\hspace*{0.3cm} {(a)} \hspace{7cm} {(b)} \hspace{2cm} \\[1mm]
\caption
Non-symmetric and symmetric extension of (a)
the metric function $f$ and (b) the dilaton function $\phi$ for
$\alpha/r_0^2=0.02$ and $f_0=1.1$ versus $l$.}
\end{figure}
However, wormhole solutions without curvature singularities can be constructed
when we extend the wormhole solutions to the second asymptotically
flat part of the manifold in a symmetric way. In this case jumps
appear in the derivatives of the metric and dilaton functions at $l=0$.
The jumps can be attributed to the presence of matter
located at the throat of the wormhole.
The corresponding junction conditions are of the form
\begin{equation}
\langle G_\mu^\nu-T_\mu^\nu \rangle = s_\mu^\nu \ , \ \ \ \
\langle \nabla^2 \phi
-\alpha \gamma e^{-\gamma \phi}R_{GB}^2 \rangle
= s_{\rm dil} \ ,
\end{equation}
where $s_\mu^\nu$ is the stress-energy tensor of the matter at the throat,
and $s_{\rm dil}$ the corresponding source term of the dilaton field.
The lhs of the junction conditions can be derived in a standard way
by integrating the Einstein and dilaton equations across the boundary
$l=0$, i.~e.
$\langle G_\mu^\nu-T_\mu^\nu \rangle
= \frac{1}{2}\lim_{L \to 0}\int_{-L}^{L}\left( G_\mu^\nu-T_\mu^\nu\right) dl$.
This yields
\begin{eqnarray}
\langle G_0^0-T_0^0 \rangle
& = &
-\frac{8\alpha \gamma e^{-\gamma \phi_0}\phi'_0}{\sqrt{f_0}r_0^2} \ ,
\label{Junc00}\\
\langle G_l^l-T_l^l \rangle
& = & 0 \ ,
\label{Juncll}\\
\langle G_\theta^\theta-T_\theta^\theta \rangle
& = & 2\frac{\nu_0'}{\sqrt{f_0}} \ ,
\label{Junctt}\\
\langle G_\vphi^\vphi-T_\vphi^\vphi \rangle
& = &
\langle G_\theta^\theta-T_\theta^\theta \rangle \ ,
\label{Juncpp}\\
\langle \nabla^2 \phi
-\alpha \gamma e^{-\gamma \phi}R_{GB}^2 \rangle
& = &
\frac{\phi_0'}{\sqrt{f_0}}
+8 \frac{\alpha\gamma e^{-\gamma \phi_0}}{\sqrt{f_0}r_0^2} \nu_0' \ ,
\label{Juncdil}
\end{eqnarray}
where the subscript $0$ indicates evaluation at $l=0$.
Next we assume that the matter at the throat takes the form of
a perfect fluid with energy density $\rho$
and pressure $p$ and a dilaton charge $\rho_{\rm dil}$.
We also introduce the action
\begin{equation}
S_\Sigma = \int\left(\lambda_1 +\lambda_0 2 \alpha e^{-\gamma \phi}\bar{R}
\right)\sqrt{-\bar{h}} d^3x
\label{S_sig}
\end{equation}
at the throat,
where $\bar{h}_{ab}$ denotes the (2+1)-dimensional induced metric on the throat,
$\bar{R}$ the corresponding Ricci scalar,
and $\lambda_1$, $\lambda_2$ are constants.
Inserting the metric
this brings the non-trivial junction conditions Eqs.~(\ref{Junc00})-(\ref{Juncdil}) to the form
\begin{eqnarray}
\frac{8 \alpha \gamma e^{-\gamma \phi_0}}{r_0^2}\,\frac{\phi_0'}{\sqrt{f_0}}
&=&
\rho-\lambda_0\,\frac{4 \alpha e^{-\gamma \phi_0}}{r_0^2}-\lambda_1
\,, \\
\frac{2\nu'_0}{\sqrt{f_0}} &=& p+\lambda_1 \,,\\
\left(\phi'_0 +\frac{8 \alpha\gamma e^{-\gamma \phi_0}}{r_0^2}\,{\nu_0'}\right)\frac{1}{\sqrt{f_0}}
&=&
\lambda_0\,\frac{4 \alpha\gamma e^{-\gamma \phi_0}}{r_0^2} +\frac{\rho_{\rm dil}}{2}\,.
\end{eqnarray}
Using these equations $\rho$, $p$ and $\rho_{\rm dil}$
can be expressed in terms of
the metric and dilaton functions and the constants $\lambda_0$, $\lambda_1$.
In Figs.~\ref{Junction-1}
we give an example for $\lambda_0=\lambda_1=1$ and $\gamma=1$.
We note that the stable wormhole solutions possess positive energy density $\rho$.
\begin{figure}[h!]
\lbfig{Junction-1}
\begin{center}
\hspace{0.0cm} \hspace{-0.6cm}
\includegraphics[height=.27\textheight, angle =0]{figD11_rho.eps}
\hspace{0.5cm} \hspace{-0.6cm}
\includegraphics[height=.27\textheight, angle =0]{figD11_p.eps}
\\[1mm]
\hspace*{0.0cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\hspace{0.0cm} \hspace{-0.6cm}
\includegraphics[height=.27\textheight, angle =0]{figD11_rhod.eps}
\hspace{0.5cm} \hspace{-0.6cm}
\includegraphics[height=.27\textheight, angle =0]{figD11_V.eps}\\
\hspace*{0.0cm} {(c)} \hspace{8cm} {(d)} \hspace{2cm} \\[1mm]
\end{center}
\caption
(a) Energy density $\rho$ , (b) pressure $p$,
(c) dilaton charge density $\rho_{\rm dil}$ and (d) interaction potential
versus $\phi_0$.}
\end{figure}
Let us also consider the special case $p=0$ (i.e.~dust)
and choose the dilaton charge density $\rho_{\rm dil}$
at the throat to be twice the dilaton charge density of the wormhole
in the `bulk', $\rho_{\rm dil} = 2\phi'_0/\sqrt{f_0}$.
This yields
\begin{eqnarray}
\lambda_1 & = & \lambda_0 = \frac{2 \nu'_0}{\sqrt{f_0}}
\label{eq_lams} \\
\rho & = & \frac{2 \nu'_0}{\sqrt{f_0}} \left(1+ \frac{4\alpha e^{-\gamma\phi_0}}{r_0^2}\right)
+\frac{8 \alpha\gamma e^{-\gamma\phi_0}}{r_0^2} \frac{\phi'_0}{\sqrt{f_0}}
\nonumber\\
& = & \frac{2 \nu'_0}{\sqrt{f_0}} + \frac{8\alpha e^{-\gamma\phi_0}}{r_0^2}
\left(\nu'_0 + \gamma\phi'_0 \right)\frac{1}{\sqrt{f_0}} \ .
\label{eq_rho2}
\end{eqnarray}
Interestingly, $\lambda_0=\lambda_1$ and $\rho$
are positive for all wormhole solutions,
as shown in Fig~\ref{Junction-3}.
\begin{figure}[h!]
\lbfig{Junction-3}
\centerline{
\includegraphics[height=.28\textheight, angle =0]{Figdust_rho.eps}
\includegraphics[height=.28\textheight, angle =0]{Figdust_lam.eps}
}
\hspace*{0.0cm} {(a)} \hspace{8cm} {(b)} \hspace{2cm} \\[1mm]
\caption
(a) Energy density $\rho$ and (b) $\lambda=\lambda_0=\lambda_1$
versus $\phi_0$
for dust.}
\end{figure}
Next we consider perturbations of the wormhole solutions in the
form $\tilde{f}(l,t) = f(l) + \delta f(l) e^{i\sigma t}$, etc. For symmetric wormholes the derivatives
of the perturbations also have a jump at $\Sigma$. Introducing perturbations of the
energy density $\delta \rho e^{i\sigma t}$, pressure $\delta p e^{i\sigma t}$ and dilaton charge
$\delta \rho_{\rm dil} e^{i\sigma t}$
we find
\begin{eqnarray}
\delta \rho & = & \frac{4\alpha e^{-\phi_0}}{r_0^2\sqrt{f_0}}
\left( \delta \phi'_0 -2\frac{\delta f_0}{f_0} \phi'_0
-\left( 2\phi'_0 +\lambda_0\sqrt{f_0} \right)\delta \phi_0\right)
\label{eq_dprho} \\
\delta p & = & 2\frac{\delta \nu'_0}{\sqrt{f_0}}
-\frac{\delta f_0}{f_0}\frac{\nu'_0}{\sqrt{f_0}}
-4\lambda_0 \alpha e^{-\phi_0}\sigma^2 e^{-2\nu_0}\delta \phi_0
\label{eq_dp} \\
\delta \rho_{\rm dil} & = & \delta \phi'_0
+\frac{8 \alpha e^{-\phi_0} }{r_0^2} \delta \nu'_0
-\frac{\delta f_0}{2 f_0}\left(\phi'_0 +\frac{8 \alpha e^{-\phi_0}}{r_0^2} \nu'_0\right)
+\frac{4\alpha e^{-\phi_0}}{r_0^2}\left(\sqrt{f_0}\lambda_0 - 2\nu'_0\right)\delta \phi_0
\label{eq_dphi}
\end{eqnarray}
Thus these equations give the perturbations of the energy density, the pressure
and the dilaton charge on the throat. Thus for any perturbation in the `bulk', we
can find the corresponding perturbation on the throat. Consequently, the stability
of the solutions is not affected by the matter on the throat.
\section{Geodesics}
The study of the orbits of test particles and light
is essential to fully understand the properties of a spacetime.
The motion of test particles in a EGBd wormhole spacetime
with metric $g_{\mu\nu}$, Eq.~(\ref{metricL}), and dilaton field $\phi$
is governed by the Lagrangian
\begin{equation}
{\cal L} = \frac{1}{2} e^{-2\beta \phi} g_{\mu\nu} \dot{x}^\mu \dot{x}^\nu \ ,
\label{lag}
\end{equation}
where the dot denotes the derivative with respect to the affine parameter $\tau$,
and $\beta$ is a constant ($\beta =1/2$ for heterotic string theory).
The conjugate momenta
$p_\mu = \frac{\partial {\cal L}}{\partial\dot{x}^\mu}$
are found to be
\begin{equation}
p_t = - e^{-2\beta \phi} e^{2\nu}\dot{t}\ , \ \ \
p_l = e^{-2\beta \phi} f\dot{l}\ , \ \ \
p_\theta = e^{-2\beta \phi} (r_0^2+l^2)\dot{\theta}\ , \ \ \
p_\vphi = e^{-2\beta \phi} (r_0^2+l^2)\sin^2\theta\dot{\vphi}\ .
\label{momenta}
\end{equation}
First integrals of motion are given by
\begin{equation}
p_t = const. = -E \ ,\ \ \ p_\vphi= const. = L \ .
\label{first_int}
\end{equation}
We refer to $E$ as the energy of the test particle
and to $L$ as its angular momentum.
The affine parameter can be chosen such that $2 {\cal L} = \hat{\kappa}$,
with $\hat{\kappa}=-1$ for time-like geodesics and $\hat{\kappa}=0$ for null geodesics.
Since we are considering spherically symmetric
spacetimes we may choose $\theta = \pi/2$, which implies $\dot{\theta}=0$.
Employing Eqs.~(\ref{first_int}) the Lagrangian then reduces to
\begin{equation}
2{\cal L} = e^{2\beta\phi}e^{-2\nu}
\left[-E^2 + e^{2\nu}\left(e^{-4\beta\phi}f \dot{l}^2
+\frac{L^2}{r_0^2+l^2}\right)\right]
= \hat{\kappa} \ .
\label{lageff}
\end{equation}
Let us first consider time-like geodesics.
For $\hat{\kappa} = -1$ we obtain
\begin{equation}
\dot{l}^2 = \frac{e^{4\beta\phi}e^{-2\nu}}{f}
\left[ E^2 -V^2_{\rm eff}(l,L)\right] \ ,
\label{dotl2}
\end{equation}
where we introduced the effective potential
\begin{eqnarray}
V^2_{\rm eff}(l,L) =
e^{2\nu}\left(-\hat{\kappa} e^{-2\beta\phi}+\frac{L^2}{r_0^2+l^2}\right)
\quad
\Longleftrightarrow
\quad
V^2_{\rm eff}(r,L) =
e^{2\nu}\left(-\hat{\kappa} e^{-2\beta\phi}+\frac{L^2}{r^2}\right) \ .
\label{veff}
\end{eqnarray}
The effective potential $V_{\rm eff}$
is a suitable quantity to discuss qualitatively
the trajectories of test particles.
We note that the effective potential
approaches the value of one asymptotically.
Thus it follows from Eq.~(\ref{dotl2})
that unbound trajectories are only possible if $E \ge 1$.
The turning points $r_i$ of trajectories for a given
energy $E$ and angular momentum $L$ are determined by the condition
\begin{equation}
E= V_{\rm eff}(r_i,L) \ .
\end{equation}
We note that, in contrast to black hole spacetimes, the wormhole
spacetimes do not possess an event horizon. Consequently, particles
cannot disappear behind an event horizon.
Instead they can travel through the throat of the wormhole
from one asymptotically flat part of the manifold to the other.
\begin{figure}[h!]
\lbfig{ff-1}
\begin{center}
\includegraphics[height=.33\textheight, angle =0]{veff1.1_all.eps}
\end{center}
\caption
Effective potential $V_{\rm eff}$ versus $\log_{10}(r/r_0)$ for several
values of the angular momentum $L$ for the wormhole solution with
$\alpha/r_0^2=0.05$ and $f_0=1.1$.
}
\end{figure}
To discuss the various types of trajectories in these
wormhole spacetimes we consider as an example
the wormhole solution with parameters $\alpha=0.05$ and $f_0=1.1$.
We exhibit the corresponding effective potential $V_{\rm eff}(r,L)$
in Fig.~\ref{ff-1} for several values
of the angular momentum $L$.
In particular, we first discuss the dependence of the shape
of the effective potential on the angular momentum $L$ of the test particle,
and next consider the different types of trajectories
with their dependence on the energy $E$ of the test particle.
In the case $L=0$
the effective potential is a monotonically increasing function of $r$,
assuming its minimum at the throat $r=r_0$ and tending to one
asymptotically. When the angular momentum is increased, the effective potential
remains monotonic until it develops a saddle point for some critical
value $L_{\rm crit}$.
For $L>L_{\rm crit}$ the effective potential is no longer monotonic,
since a local maximum $V_{\rm max}(L)$
and a local minimum $V_{\rm min}(L)$ occur.
If $L$ is larger than a certain value, $L_1$ (say),
the maximum $V_{\rm max}(L)$ exceeds the asymptotic value of the
effective potential, and thus it becomes the global maximum.
We note that the essential features of the
effective potential $V_{\rm eff}(r,L)$ are
the same for all the wormhole solutions considered.
Concerning the types of orbits of test particles we note that there are
two kinds of trajectories.
For trajectories of the first kind the particles remain on a single
asymptotically flat manifold, whereas for trajectories of
the second kind the particles travel from one asymptotically flat part
of the manifold to the other,
passing through the throat.
Trajectories of the first kind exist only if the effective potential
is non-monotonic, i.e.~for $L>L_{\rm crit}$.
Bound orbits exist for $V_{\rm min}(L)\leq E<1$,
e.g.~$E_2$ in Fig.~\ref{ff-1},
whereas unbound orbits exist for $1< E <V_{\rm max}(L)$,
e.g.~$E_3$ in Fig.~\ref{ff-1}.
For unbound orbits we need $L>L_1$,
thus $V_{\rm max}(L)$ must be the absolute maximum.
Trajectories of the second kind exist for all values
of $E$ above the minimum of the effective potential
$V_{\rm eff}(r_0,L)$.
The particles move on bound orbits if
$V_{\rm eff}(r_0,L)\leq E < \max\{1,V_{\rm max}(L)\}$,
e.g.~$E_1$, $E_2$, $E_3$ in Fig.~\ref{ff-1}.
Thus for $E=V_{\rm eff}(r_0,L)$ the particles move on circles at the
throat. Radial trajectories are obtained for $L=0$.
In this case the particles oscillate about the throat.
In the limit $E_0=\sqrt{V_{\rm eff}(r_0,0)}$
they possess the minimal possible energy
and are at rest at the throat.
On the other hand, if $E>\max\{1,V_{\rm max}(L)\}$,
e.g.~$E_4$ in Fig.~\ref{ff-1},
the particles travel along unbound orbits, starting at infinity on one
asymptotically flat part of the manifold, passing through the throat and reaching
infinity on the other asymptotically flat part of the manifold.
In Fig.~\ref{ff-2} we exhibit the minimal possible energy
$E_0=\sqrt{V_{\rm eff}(r_0,0)}$ versus $D/M$ for several values of
$\alpha/r_0^2$. Note that $E_0$ approaches zero in the black hole limit.
The black line on the left, highlighted in the inset,
indicates the stability change of the wormhole solutions.
\begin{figure}[h!]
\lbfig{ff-2}
\begin{center}
\includegraphics[height=.25\textheight, angle =0]{FiglE.eps}
\end{center}
\caption
Minimal energy $E_0=\sqrt{V_{\rm eff}(r_0,0)}$ of test particles
versus the scaled dilaton charge $D/M$
for several values of $\alpha/r_0^2$.
The black line on the left denoted by $\sigma^2=0$
indicates the stability change of the wormhole solutions.
}
\end{figure}
In order to calculate the trajectories we consider
\begin{equation}
\frac{d\vphi}{dl} = \frac{\dot{\vphi}}{\dot{l}} \nonumber \\
= \pm L \frac{e^\nu}{r_0^2+l^2}\sqrt{\frac{f}{E^2 -V^2_{\rm eff}}} \ ,
\nonumber
\end{equation}
from which we find
\begin{equation}
\vphi(l) = \vphi_0
\pm L \int_{l_0}^l \frac{e^\nu}{r_0^2+l^2}\sqrt{\frac{f}{E^2 -V^2_{\rm eff}}} dl' \ ,
\end{equation}
where $l$ is restricted to intervals where $E^2 -V^2_{\rm eff} \geq 0$ and
$\vphi_0$ is an integration constant.
The trajectories are then displayed in the $xy$-plane, where
\begin{equation}
x(l) = \sqrt{r_0^2+l^2}\cos(\vphi(l)) \ , \ \ \ \
y(l) = \sqrt{r_0^2+l^2}\sin(\vphi(l)) \ .
\end{equation}
In Fig.~\ref{ff-3} we show examples of bound orbits
of test particles with angular momentum $L=2$
and several values of the energy $E$
for the wormhole solution with $\alpha/r_0^2=0.05$ and $f_0=1.1$.
For an energy of $E=E_1=0.9$ (Fig.~\ref{ff-3}a)
there exists only a bound orbit of the second kind.
This is highlighted in the figure by using different colors (line styles) for
the two asymptotically flat manifolds, when projecting into the
$xy$-plane.
In contrast, for $E=E_2=0.98$ (Fig.~\ref{ff-3}b)
there exists in addition a bound orbit of the first kind.
Examples for unbound orbits are shown in Fig.~\ref{ff-4}a for
energies $E=E_3=1.02$ and $E=E_4=1.05$.
\begin{figure}[h!]
\lbfig{ff-3}
\begin{center}
a) \includegraphics[height=.25\textheight, angle =0]{orb1.1S_E0.9.eps}
\includegraphics[height=.25\textheight, angle =0]{orb1.1S3d_E0.9.eps}\\
b) \includegraphics[height=.25\textheight, angle =0]{orb1.1S_E0.98.eps}
\includegraphics[height=.25\textheight, angle =0]{orb1.1L_E0.98.eps}
\end{center}
\caption
Bound orbits of massive test particles with angular momentum $L=2$
and energy (a) $E=0.9$ and (b) $E=0.98$
in the wormhole spacetime with $\alpha/r_0^2=0.05$ and $f_0=1.1$.
The colours red (solid) and blue (dotted) indicate
motion on the first and the second asymptotically flat part of
the manifold, respectively.
The throat of the wormhole is shown by the black circle.
The right panel of a) shows the orbit on the isometric embedding of the wormhole.
}
\end{figure}
\begin{figure}[h!]
\lbfig{ff-4}
\begin{center}
\includegraphics[height=.25\textheight, angle =0]{orbU1.1.eps}
\includegraphics[height=.25\textheight, angle =0]{veff1.1_N.eps}
\end{center}
\hspace*{0.7cm} {(a)} \hspace{7.5cm} {(b)} \hspace{2cm} \\[1mm]
\caption
(a) Same as Fig.~\ref{ff-3} for unbound orbits of massive test particles
and energies $E=1.02$, $1.05$.
(b) ``Normalized'' effective potential $v_{\rm eff}$ for massless test particles
versus $\log_{10}(r/r_0)$ for the wormhole solution with $\alpha/r_0^2=0.05$ and $f_0=1.1$.
}
\end{figure}
Let us now consider the null geodesics in these wormhole spacetimes.
Setting $\hat{\kappa}=0$ in Eq.~(\ref{lageff}) we find
\begin{equation}
\dot{l}^2 = \frac{e^{4\beta\phi}e^{-2\nu}}{f}
\left[ E^2 -L^2\frac{e^{2\nu}}{r_0^2+l^2}\right]
=E^2 \frac{e^{4\beta\phi}e^{-2\nu}}{f}
\left[1 -\left(\frac{L}{E}\right)^2 \frac{e^{2\nu}}{r_0^2+l^2}\right]\ .
\label{dotl2N}
\end{equation}
We define an effective potential $V_{\rm eff}(l,L/E)$ by
\begin{equation}
V^2_{\rm eff}(l,L/E) =
\left(\frac{L}{E}\right)^2 \frac{e^{2\nu}}{r_0^2+l^2}
= \left(\frac{L}{E}\right)^2 v^2_{\rm eff}(l)\ .
\label{veffN}
\end{equation}
We note that for the discussion of the null geodesics it is more convenient
to consider the ``normalized'' effective potential
$v_{\rm eff}(l)=\frac{e^{\nu}}{\sqrt{r_0^2+l^2}}$, since it is independent
of $L/E$.
As an example, we show the ``normalized'' effective
potential $v_{\rm eff}$ for the parameter values $\alpha/r_0^2 =0.05$
and $f_0=1.1$ in Fig.~\ref{ff-4}.
The ``normalized'' effective potential possesses a
local minimum $v_{\rm min}$ at the throat,
a maximum $v_{\rm max}$ at some distance from the throat
and tends to zero asymptotically.
As for the timelike geodesics there are two kinds of trajectories:
either the massless test particle remains on one of the asymptotically
flat parts of the manifold,
or it passes through the throat from one asymptotically flat part to the other.
The first kind of (unbound) trajectories exists only if the ratio $E/L$ is
smaller than the maximum of the ``normalized'' effective potential.
The second kind of trajectories includes unbound geodesics if
$E/L > v_{\rm max}$ and bound geodesics if $v_{\rm min}\leq E/L < v_{\rm max}$.
Circular orbits exist for $E/L=v_{\rm min}=v_{\rm eff}(l=0)$
(and unstable ones for $E/L=v_{\rm max}$).
\section{Acceleration and Tidal Forces}
Following the formalism of \cite{Morris:1988cz}, we will now calculate the magnitude of the
acceleration and tidal forces that a traveler traversing the wormhole would feel.
For this purpose, it is particularly convenient to make two changes of reference
frames: first, starting from the standard reference frame with basis vectors
$({\mathbf e}_t, {\mathbf e}_l, {\mathbf e}_\theta, {\mathbf e}_\varphi)$,
in which the line-element takes the form of Eq.~(\ref{metricL}),
we change to an orthonormal reference frame with
\begin{equation}
{\mathbf e}_{\hat t}=e^{-\nu} {\mathbf e}_t, \quad
{\mathbf e}_{\hat l}=\frac{1}{\sqrt{f}}\,{\mathbf e}_l, \quad
{\mathbf e}_{\hat \theta}=\frac{1}{r}\,{\mathbf e}_\theta, \quad
{\mathbf e}_{\hat \varphi}=\frac{1}{r \sin\theta}\,{\mathbf e}_\varphi
\label{ortho1}
\end{equation}
in terms of which the metric tensor assumes the form:
$g_{\hat \alpha \hat \beta}={\mathbf e}_{\hat \alpha} \cdot {\mathbf e}_{\hat \beta}=
\eta_{\hat \alpha \hat \beta}$.
Alternatively, we may write the set of equations in (\ref{ortho1}) as
${\mathbf e}_{\hat \alpha} =L_{\hat \alpha}^{\,\,\mu}\,{\mathbf e}_\mu$, with
\begin{equation}
L_{\hat \alpha}^{\,\,\mu}=\left[\begin{array}{rccc}
e^{-\nu} & 0 & 0 & 0 \\ 0 & 1/\sqrt{f} & 0 & 0\\ 0 & 0 & 1/r & 0\\
0 & 0 & 0 & 1/(r \sin\theta)\end{array} \right]. \label{trans1}
\end{equation}
The above allows us to write the transformation law of the components of the
Riemann tensor as
\begin{equation}
R^{\hat \alpha}_{\,\,\hat \beta \hat \gamma \hat \delta}=L^{\,\,\hat \alpha}_{\mu}
\,L_{\hat \beta}^{\,\,\nu}\,L_{\hat \gamma}^{\,\,\rho}\,L_{\hat \delta}^{\,\,\sigma}\,
R^\mu_{\,\,\nu \rho \sigma}\,,
\label{Riemann-trans1}
\end{equation}
with $L^{\,\,\hat \alpha}_\mu=(L_{\hat \alpha}^{\,\,\mu})^{-1}$. Next, we introduce
the orthonormal reference frame of the traveler which is related to the previous one
by a Lorentz transformation
\begin{equation}
{\mathbf e}_{\tilde t}=\gamma\,{\mathbf e}_{\hat t} \mp \gamma\,\frac{v}{c}\,
{\mathbf e}_{\hat l}\,, \quad
{\mathbf e}_{\tilde l}=\mp\gamma\,{\mathbf e}_{\hat l}+
\gamma\,\frac{v}{c}\,{\mathbf e}_{\hat t}\,, \quad
{\mathbf e}_{\tilde \theta}={\mathbf e}_{\hat \theta}, \quad
{\mathbf e}_{\tilde \varphi}={\mathbf e}_{\hat \varphi}\,,
\label{ortho2}
\end{equation}
with $\gamma=[1-(v/c)^2)]^{-1/2}$ and $v=\mp (\sqrt{f} dl/e^\nu dt)$ the
radial velocity of the traveler at radius $l$ as measured by a static
observer there. As before, we may write
${\mathbf e}_{\tilde \alpha} =\Lambda_{\tilde \alpha}^{\,\,\hat \mu}\,
{\mathbf e}_{\hat \mu}$, with
\begin{equation}
\Lambda_{\tilde \alpha}^{\,\,\hat \mu}=\left[\begin{array}{cccc}
\gamma & \mp\gamma (v/c) & 0 & 0 \\ \gamma\,(v/c) & \mp \gamma & 0 & 0\\
0 & 0 & 1 & 0\\ 0 & 0 & 0 & 1\end{array} \right],
\label{trans2}
\end{equation}
so that
\begin{equation}
R^{\tilde \alpha}_{\,\,\tilde \beta \tilde \gamma \tilde \delta}=
\Lambda^{\,\,\tilde \alpha}_{\hat \mu}\,\Lambda_{\tilde \beta}^{\,\,\hat \nu}\,
\Lambda_{\tilde \gamma}^{\,\,\hat\rho}\,\Lambda_{\tilde \delta}^{\,\,\hat\sigma}\,
R^{\hat \mu}_{\,\,\hat \nu \hat \rho \hat \sigma}\,.
\label{Riemann-trans2}
\end{equation}
We note that ${\mathbf e}_{\tilde t} \cdot {\mathbf e}_{\tilde t}=-1$, and thus
${\mathbf e}_{\tilde t}$ can be naturally considered as the traveler's normalized
vector of four-velocity ${\mathbf u}$. For the magnitude of ${\mathbf u}$ to
be fixed in the rest frame of the traveler, the four-acceleration
${\mathbf a}=d{\mathbf u}/d \tau$ should be orthogonal to the four-velocity,
i.e.~${\mathbf a} \cdot {\mathbf u}=0$, and therefore ${\mathbf a}=(0, a^i)$ --
if we further assume that the traveler moves radially, then
${\mathbf a}= a {\mathbf e}_{\tilde l}$. For a traveler moving along their
world-line in which the tangent vector is
${\mathbf e}_{\tilde t}={\mathbf u}$, the acceleration is given by the formula
${\mathbf a}={\mathbf \nabla}_{\mathbf u}{\mathbf u}$ or
$a^{\tilde \mu}=u^{\tilde \mu}_{;\tilde a} u^{\tilde a} c^2$.
In order to compute the magnitude of the acceleration that the traveler feels,
we work in the following way: we calculate the $a_t$ component in the
$(t,l,\theta,\varphi)$ coordinate frame first according to the formula
\begin{equation}
\frac{a_t}{c^2}=u_{t;a}u^a=u_{t,a} u^a - \Gamma^\lambda_{ta}u_\lambda u^a
=u_{t,l} u^l-\Gamma^t_{tl} u_t u^l-\Gamma^l_{tt}u_lu^t=
\pm\gamma\,\frac{v}{c}\,\frac{1}{\sqrt{f}}\,(e^\nu \gamma)'\,,
\end{equation}
where we have assumed that the four-velocity is a function of the radial variable
$l$ and rewritten its expression as
\begin{equation}
{\mathbf u}=\gamma e^{-\nu} {\mathbf e}_t
\mp \gamma\,\frac{v}{c}\,\frac{1}{\sqrt{f}}\,{\mathbf e}_l \equiv u^t {\mathbf e}_t
+u^l\,{\mathbf e}_l\,.
\end{equation}
However, it also holds that
\begin{equation}
a_t={\mathbf a} \cdot {\mathbf e}_{t}=(a {\mathbf e}_{\tilde l}) \cdot {\mathbf e}_{t}
= a \gamma \frac{v}{c}\,e^{-\nu} ({\mathbf e}_t \cdot {\mathbf e}_t)=
-a \gamma \frac{v}{c}\,e^{\nu}\,.
\end{equation}
Combining the above two results for $a_t$, we find that the magnitude of the
acceleration is
\begin{equation}
|a|=c^2 \Bigl |e^{-\nu} \frac{1}{\sqrt{f}}\,(e^\nu \gamma)'\Bigr|\,.
\label{con-acc}
\end{equation}
For the wormhole to be traversable, the above quantity must remain always finite
and, if possible, take a small value.
In \cite{Morris:1988cz}, it was also demanded that the gravitational acceleration
should be small at the location of the stations where the trip starts and ends.
Demanding that the time needed to complete the whole trip should not be too large,
the stations should be fairly close to the throat. Both of these conditions are
satisfied in our case, since an asymptotically flat regime is reached fairly quickly.
The above also means that the traveler, starting from the stations, does not need
to travel with a relativistic velocity in order to approach the wormhole in a
reasonable time. Then, if $v \ll c$, $\gamma \simeq 1$, and the acceleration
(\ref{con-acc}) reduces to
\begin{equation}
|a|= c^2 \frac{|\nu'|}{\sqrt{f}}\,. \label{accel-final}
\end{equation}
In Fig. \ref{fig_acc}, we depict the dimensionless acceleration
$\hat{a}=|a|/(c^2/r_0)$ at the throat as a function of $D/M$. We observe that
$\hat{a}$ ranges roughly between 10 and 100 for a set of wormhole
solutions with parameter $\alpha/r_0^2$ ranging between
0.128 and 0.01. In the context of superstring theory, $\alpha \sim \ell_P^2$,
therefore, for the above solutions $r_0 \sim 10\,\ell_P$ and the magnitude
of the acceleration turns out to be $~(10^{51}-10^{52})\,g_\oplus$, where
$g_\oplus$ is the acceleration of gravity at the surface of the earth.
Since there is no upper limit for $r_0$ in our solutions, one may be tempted
to increase the size of the throat of the wormhole so that $|a|$ is of the order
of $g_\oplus$; this would demand a throat radius of the order of
at least $(10-100)$ light-years. The aforementioned result could perhaps have
been anticipated since our analysis is performed within the context of
superstring theory, a theory whose fundamental scale is tied to the Planck scale.
\begin{figure}[ht]
\lbfig{fig_acc}
\begin{center}
\includegraphics[height=.25\textheight, angle =0]{fig_acc.eps}
\end{center}
\caption
The dimensionless acceleration $\hat{a}=|a|/(c^2/r_0)$ that a traveler would
feel traversing the wormhole as a function of $D/M$ for a variety of values
of $\alpha/r_0^2$.
}
\end{figure}
Next we turn to the tidal acceleration that a traveler feels between two parts
of her body as the wormhole is crossed. Let ${\mathbf w}$ be the vector separation
between these two parts, which in the reference frame of the traveler is purely
spatial, ${\mathbf w}=(0,w^i)$. Then, the tidal acceleration is given by the expression
\cite{Morris:1988cz}
\begin{equation}
\Delta a^{\tilde \mu}=-c^2 R^{\tilde \mu}_{\,\,\tilde \nu \tilde \rho \tilde\sigma}\,
u^{\tilde \nu} w^{\tilde \rho}\,u^{\tilde \sigma}=
-c^2 R^{\tilde \mu}_{\,\,\tilde t \tilde \kappa \tilde t}\,w^{\tilde \kappa}\,,
\label{tidal}
\end{equation}
where ${\tilde \kappa}$ takes on only spatial values and where we have used that
$u^{\tilde a}=\delta^{\tilde a}_{\tilde t}$. But since the metric is diagonal and the
Riemann tensor is antisymmetric in the first two indices, the superscript ${\tilde \mu}$
should also take only spatial values. Then, the non-vanishing components of the
tidal acceleration are:
\begin{equation}
\Delta a^{\tilde l}=-c^2 R^{\tilde l}_{\,\,\tilde t \tilde l \tilde t}\,w^{\tilde l}\,,
\qquad
\Delta a^{\tilde \theta}=-c^2 R^{\tilde \theta}_{\,\,\tilde t \tilde \theta \tilde t}\,
w^{\tilde \theta}\,,
\qquad
\Delta a^{\tilde \varphi}=-c^2 R^{\tilde \varphi}_{\,\,\tilde t \tilde \varphi \tilde t}\,
w^{\tilde \varphi}\,.
\label{tidal-2}
\end{equation}
The transformation laws (\ref{trans1}) and (\ref{trans2}) allow us to compute the
components of the Riemann tensor in the orthonormal frame of the moving observer
in terms of the ones in the $(t,l,\theta,\varphi)$ coordinates. We then obtain
\begin{eqnarray}
\Delta a^{\tilde l}&=& c^2 w^{\tilde l}\,R^{\hat t}_{\,\,\hat l \hat t \hat l}
= c^2 w^{\tilde l}\,\frac{1}{f}\,R^{t}_{\,\,l t l}=
c^2\,w^{\tilde l}\,\frac{1}{f}\left(\frac{f'\nu'}{2f}-\nu'^2-\nu''\right)\,.
\end{eqnarray}
Similarly, we find:
\begin{eqnarray}
\Delta a^{\tilde \theta}=\Delta a^{\tilde \varphi} &=&
-c^2 w^{\tilde \theta}
\left(\gamma^2\,R^{\hat \theta}_{\,\,\hat t \hat \theta \hat t}+
\gamma^2\,\frac{v^2}{c^2}\,
R^{\hat \theta}_{\,\,\hat l \hat \theta \hat l}\right) =
-c^2\,w^{\tilde \theta}\,\gamma^2\left(e^{-2\nu}
R^{\theta}_{\,\, t \theta t} + \frac{v^2}{c^2 f}\,
R^{\theta}_{\,\, l \theta l}\right) \nonumber \\
&=&
-\frac{c^2\,w^{\tilde \theta}\,\gamma^2}{r_0^2+l^2}
\left[\frac{l\nu'}{f}+\frac{v^2}{c^2 f}\,\left(\frac{l f'}{2f}-
\frac{r_0^2}{r_0^2+l^2}\right)\right]\,.
\end{eqnarray}
As in the case of the acceleration, we should demand that the magnitude of the
above components should be small, i.e. of the order of the acceleration of gravity
at the surface of the earth. Using the fact that $|{\mathbf w}| \simeq 2\,{\rm m}$,
the above constraints may be written as
\begin{eqnarray}
\frac{1}{f}\left|\frac{f'\nu'}{2f}-\nu'^2-\nu''\right|
&\leq & \frac{g_\oplus}{c^2 (2\,{\rm m})}=\frac{1}{(10^{10} {\rm cm})^2}\,,
\\[2mm]
\frac{\gamma^2}{r_0^2+l^2}
\left|\frac{l\nu'}{f}+\frac{v^2}{c^2 f}\,\left(\frac{l f'}{2f}-
\frac{r_0^2}{r_0^2+l^2}\right)\right|
&\leq & \frac{g_\oplus}{c^2 (2\,{\rm m})}=\frac{1}{(10^{10} {\rm cm})^2}\,.
\end{eqnarray}
The second inequality involves the velocity with which the traveler moves and thus
may be considered as a constraint on this quantity. The first inequality restricts
again the profile of the metric functions - a similar analysis to the one above leads to
the same results regarding the magnitude of the tidal forces and the necessary
size of the throat in order to bring these down to a reasonable value.
\section{Conclusions}
The existence of traversable wormholes in the context of General Relativity
relies on the presence of some form of exotic matter. However, in the
framework of a string-inspired generalized theory of gravity, the situation
may be completely different.
Here we have investigated wormhole solutions in EGBd theory,
which corresponds to a simplified action
that is motivated by the low-energy heterotic string theory.
Indeed, as we have demonstrated, EGBd theory allows for stable,
traversable wormhole solutions, without the need of introducing
any form of exotic matter. The violation of the energy conditions,
that is essential for the existence of the wormhole solutions
\footnote{According to recent studies \cite{Houndjo},
wormhole solutions arise also in the context of $f(T)$ gravitational
theories, where $T$ is the torsion, without the energy conditions being violated.},
is realized via the presence of an effective energy-momentum tensor
generated by the quadratic-in-curvature Gauss-Bonnet term.
We have determined the domain of existence of these wormhole solutions
and shown that it is bounded by three sets of limiting solutions.
The first boundary consists of the EGBd black hole solutions
of \cite{Kanti:1995vq}.
The second boundary is approached asymptotically,
when the curvature radius at the throat of the wormhole diverges
(the $f_0 \to \infty$ limit).
Finally, at the third boundary, solutions with a curvature singularity
at a finite distance from the throat are encountered.
We have investigated the properties of these EGBd wormholes
and derived a Smarr-like mass relation for them. In this,
the horizon properties in the black hole case
are replaced by the corresponding throat properties of the wormholes,
thus the area and surface gravity here refer to the ones at the throat.
Moreover, as is well-known for black holes in EGB theories,
their entropy does not correspond merely to their horizon area but it
receives a GB correction term. Similarly, we find that the
mass formula for the wormhole solutions includes an analogous GB
correction term; in addition, another term, that vanishes in the
black hole case, appears that represents the GB corrected dilaton
charge at the throat. We have demonstrated that the Smarr relation is
satisfied very well by the numerical solutions.
We have also investigated the stability of the solutions.
We have shown that a subset of our wormhole solutions, the one that
lies close to the border with the linearly stable dilatonic black holes
in the domain of existence, is also linearly stable
with respect to radial perturbations.
While we have also shown that another subset is unstable,
we have concluded that the study of the standard equivalent Schr\"odinger equation
cannot determine the stability for the full domain of existence
We hope to resume the question of the existence of an alternative
method for the study of the stability of all of our wormhole
solutions at some future work.
When the wormhole solutions are extended to the second
asymptotically flat region,
this extension must be made in a symmetric way,
since otherwise a singularity is encountered.
As a consequence of this symmetric extension,
the derivatives of the metric and dilaton functions
become discontinuous at the throat. This discontinuity
demands the introduction of some matter distribution
at the throat. We have shown that this may be realized
by the introduction of a perfect fluid at the throat whose
energy density is positive for the subset of stable wormhole solutions.
Next, we have studied and classified the geodesics
of massive and massless test particles in these wormhole spacetimes.
Depending on the respective effective potential,
there are two general kinds of trajectories for these particles.
The particles may remain on bound or escape orbits
within a single asymptotically flat part of the spacetime,
or they may travel from one asymptotically flat part
to the other on escape orbits, and travel back and forth
on bound orbits.
In addition, we have calculated the acceleration and tidal forces
which travelers traversing the wormhole would feel. We find that
their magnitude may be small for fairly large values of the size
of the throat of the wormhole. According to our findings, the radius
of the wormhole throat is bounded from below only, therefore, the
wormholes can be indeed arbitrarily large. Astrophysical consequences
will be addressed in a forthcoming paper as well as the existence
of stationary rotating wormhole solutions in the EGBd theory.
\section*{Acknowledgments}
We gratefully acknowledge discussions with Eugen Radu.
B.K. acknowledges support by the DFG.
|
\section{Introduction}
Control charts such as the Shewhart chart \citep{Shewhart1931ECo} and
the cumulative sum (CUSUM) chart \citep{Page1954CIS} have been
valuable tools in many areas, including reliability
\citep{OConnor2002Pre,Xie2002Sec}, medicine
\citep{Carey2003Ihw,Lawson2005Sas,Woodall2006Tuo} and finance
\citep{Frisen2008FS}. See \cite{Stoumbos2000SoS} and the special
issues of ``Sequential Analysis'' (2007, Volume 26, Issues 2,3) for an
overview. Often, heterogeneity between observations is accounted for
by using risk-adjusted charts based on fitted regression models
\citep{Grigg2004oor,Horvath2004Mci,Gandy2010ram}.
A common convention in monitoring based on control charts is to assume
the probability distribution of in-control data to be known. In
practice this usually means that the distribution is estimated based
on a sample of in-control data and the estimation error is ignored.
Examples of this are
\cite{Steiner2000Msp,Grigg2004oor,Bottle2008Iin,Biswas2008rCi,Fouladirad2008Otu,Sego2009Rmo,Gandy2010ram}.
However, the estimation error has a profound effect on the performance
of control charts. This has been mentioned at several places in the
literature, e.g.\ \cite{jones2004rld,Albers2004Esc,jensen2006epe,Stoumbos2000SoS,Champ2007PoM}.
To illustrate the effect of estimation, we consider a CUSUM chart
\citep{Page1954CIS} with normal observations and estimated in-control
mean. We observe a stream of independent random variables
$X_1,X_2,\ldots$ which in control have an $ N(\mu,1)$ distribution and
out of control have an $N(\mu+\Delta,1)$ distribution, where $\Delta>0$
is the shift in the mean. The chart
switches from the in-control state to the out-of-control state at an
unknown time $\kappa$. The unknown in-control mean $\mu$ is estimated
by the average $\hat\mu$ of $n$ past in-control observations
$X_{-n},\dots,X_{-1}$ (this is often called phase 1 of the monitoring;
the running of the chart is called phase 2). We consider the CUSUM chart
$$S_t=\max(0, S_{t-1}+X_t-\hat \mu - \Delta/2), \quad S_0=0 $$
with hitting time $\tau=\inf\{t>0: S_t\geq c\}$ for some threshold
$c>0$.
\begin{figure}[tb]
\centering
\includegraphics[width=0.95\linewidth]{simpaper/estimerr_guaranteed_CUSUM.pdf}
\caption{In-control distribution of ARL=$\E(\tau|\hat\mu)$ for
CUSUMs for standard normally distributed data. The mean $\hat
\mu$ used in the monitoring is estimated based on $n$ past
observations. The boxplots show the 2.5\%, 10\%, 25\%, 50\%,
75\%, 90\% and 97.5\% quantiles.The top part of the plot shows
the situation when estimation error is ignored. In the middle part the
threshold has been chosen to give an unconditional ARL of 100
(averaging out the parameter estimation). In the bottom part the threshold is
adjusted to guarantee with 90\% probability an in-control ARL of at least 100. }
\label{fig:estimerr}
\end{figure}
The in-control average run length,
$\ARL=\E(\tau|\hat\mu,\kappa=\infty)$, depends on $\hat \mu$ and is
thus a random quantity. The top part of the plot in
Figure~\ref{fig:estimerr} shows boxplots of its distributions with
threshold $c=2.84$, $\Delta=1$ and various numbers of past
observations. If $\hat \mu=\mu$, i.e.\ $\mu$ was know, this
would give an in-control $\ARL$ of 100. The estimation error is having a substantial effect on
the attained $\ARL$ even for large samples such as
$n=1000$.
For further illustrations of the impact of estimation error see
\cite{jones2004rld} for CUSUM charts and \cite{Albers2004Esc} for
Shewhart charts.
So far, no general approach for taking the estimation error into
account has been developed, but there are many special constructions
for specific situations. For instance, for some charts so called
self-starting charts
\citep{Hawkins1987SCC,Hawkins1998csc,Sullivan2002SCC}, maximum
likelihood surveillance statistics
to eliminate parameters
\cite[e.g.][]{Frisen2009}, correction
factors for thresholds
\citep{Albers2004Esc,jones2002statistical},
modified thresholds \citep{Zhang2011TSX} and threshold
functions \citep{Horvath2004Mci,Aue2006Cpm} have been
developed. Various bootstrap schemes for specific situations have also
been suggested, see for instance
\cite{Kirch2008BSC,Chatterjee2009Dcs,Capizzi2009Bdo,Huskova10Bsc}.
Further, some nonparametric charts which account for the estimation
error in past data have been proposed, see \cite{Chakraborti2007Ncc}
and references therein. Recently some modified charts for monitoring
variance in the normal distribution with estimated parameters have
been suggested by \cite{Maravelakis2009AEC} and
\cite{Castagliola2011ACC}.
When addressing estimation error, the above methods mainly focus
on the performance of the charts averaged over both the estimation of
the in-control state as well as running the chart once.
In the middle part of Figure~\ref{fig:estimerr}, the threshold has
been chosen such that, averaged over both the estimation of
the in-control state as well as running the chart once, the average
run length is $100$ (this results in a different threshold
for each $n$). It turns out that only a small change in the threshold
is needed and that the distribution of the conditional $\ARL=\E(\tau|\hat
\mu)$ is only changed slightly. This bias correction
for the $\ARL$ actually goes in the wrong direction in the sense that
it implies more short $\ARL$s. This is due to the $\ARL$ being
substantially influenced by the right tail of the run length
distribution, see the discussion in Section 2 of \cite{Albers2006SAC}.
However, usually, after the chart parameters are estimated, the
chart is run for some time without any reestimation of the in-control
state even if the chart signals. Moreover, in some situations, several
charts are run based on the same estimated parameters.
In these situations the ARL conditional on the estimated in-control
state is more relevant than the unconditional ARL. In the middle and
upper part of Figure \ref{fig:estimerr}, one sees that the conditional
ARL can be much lower than 100, meaning that both the unadjusted
threshold and the
threshold adjusted for bias in the unconditional ARL
lead, with a substantial probability, to charts
that have a considerably decreased time until false alarms.
To overcome these problems we will look at the performance of the
chart conditional on the estimated in-control distribution, averaging
only over different runs of the chart. This will lead to the
construction of charts that with high probability have an in-control
distribution with desired properties conditional on the observed past
data, thus reducing the situations in which there are many false
alarms due to estimation error.
The bottom part of Figure \ref{fig:estimerr} shows the distribution of the in control
ARL when the threshold for each set of past data is adjusted to
guarantee an in-control ARL of at least 100 with probability 90\%. The adjustment is
calculated using a bootstrap procedure explained later in the paper.
The adjustment succeeds to avoid the too low ARLs with the
prescribed probability, and we will see later that the cost in a
higher out-of-control ARL is modest. Using hitting probabilities instead of ARL as
criterion leads to similar results.
\xx{talk about adjustments}
Our approach is similar in spirit to the exceedance probability
concept developed by Albers and Kallenberg for various types of
Shewhart \citep{Albers2004AEC,Albers2005Ncf,Albers2005EPF} and
negative binomial charts \citep{Albers2009CUM,Albers2010Toc}. They
calculate approximate adjusted thresholds such that there is only a
small prescribed probability that some performance measure, for
instance an ARL, will be a certain amount below or above a specified
target.
The main difference between their approach and what we present is that
our approach applies far more widely, to many different types of
charts and without having to derive specific approximation formulas in
each setting. If we apply a nonparametric bootstrap, the
proposed procedure will be robust against model misspecification. In
addition to that, our approach allows not only to adjust the threshold
but also to give a confidence interval for the in-control performance
of a chart for a fixed threshold. Lastly, even though not strongly
advocated in this paper, the bootstrap procedure we propose can also
be used to do a bias correction for the unconditional performance of
the chart, as in the middle part of Figure \ref{fig:estimerr}.
Next, we describe our approach more formally.
Suppose we want to use a monitoring scheme and that the in-control
distribution $P$ of the observations is unknown, but that
based on past in-control behaviour we have an estimate $\hat P$ of the
in-control distribution. Let $q$ denote the in-control property of
the chart we want to compute, such as the $\ARL$, the false alarm
probability or the threshold needed for a certain $\ARL$ or false alarm
probability. In the above example we were interested to find a
threshold such that the in-control ARL is 100.
Generally, $q$ may depend on both the true in-control distribution $P$
and on estimated parameters of this distribution which for many charts
are needed to run the chart. We denote these parameters by $\hat
\xi=\xi(\hat P)$.
In the above CUSUM chart example $\hat \xi = \hat
\mu$. We are interested in $q(P;\hat\xi)$, that
is the in-control performance of the chart conditional on the
estimated parameter. In the above CUSUM example, $q(P;\hat\xi)$ is the threshold
needed to give an $\ARL$ of 100 if the observations are from the true
in-control distribution $P$ and the
estimated parameter $\hat \mu$ is used. As $P$ is not observed $q(P;\hat
\xi)$ is not observable. As mentioned above, many papers pretend
that the estimated in-control distribution $\hat P$ equals the true
in-control distribution $P$ and thus use $ q(\hat P;\hat \xi)$.
Our suggestion is to use bootstrapping of past data to construct an
approximate one-sided confidence intervals for $q(P;\hat \xi)$. From
this we get a guaranteed conditional performance of the control
scheme.
In Section~\ref{sec:monitorhomobs} we present the general idea in the
setting with homogeneous observations, and discuss this for Shewhart
and CUSUM charts. The main theoretical results are presented in
Section~\ref{sec:gentheor}, with most of the proofs given in the
Appendix. Section \ref{sec:simulsingle} contains simulations
illustrating the performance of charts for homogeneous observations.
In Section~\ref{sec:regmod} extensions to charts based on regression
and survival analysis models are presented. Some concluding comments
are given in Section~\ref{sec:conclusion}. The suggested methods are
implemented in a flexible R-package, that will be made available on the
Comprehensive R Archive Network (CRAN).
\section{Monitoring homogeneous observations}
\label{sec:monitorhomobs}
\subsection{General idea}
\label{subsec:generalidea}
Suppose that in control we have independent observations
$X_1,X_2,\dots$ following an unknown distribution $P$. We want to use
some monitoring scheme/control chart that detects when $X_{i}$ is no
longer coming from $P$. The particular examples we discuss in
this paper are Shewhart and CUSUM charts, but the methodology we
suggest applies more widely.
To run the charts, one often needs certain parameters $\xi$. For
example, in the CUSUM control chart of the introduction, we
needed $
\xi= \mu$, the assumed in-control mean. These parameters will usually
be estimated.
Let $\tau$ denote the time at which the chart signals a change. As $\tau$ may
depend on $\xi$, we sometimes write $\tau(\xi)$. The
charts we consider use a threshold $c$, which determines how quickly
the chart signals (larger $c$ lead to a later signal).
The performance of such a control chart with the in-control
distribution $P$ and the parameters $ \xi$ can, for example, be
expressed as one of the following.
\begin{itemize}%
\setlength{\itemsep}{0pt}%
\setlength{\parskip}{0pt}%
\item $\ARL(P;\xi)=\E(\tau( \xi))$, where $\E$ is the expectation with respect to $P$.
\item ${\hit}(P;\xi)=\Prob(\tau( \xi)\leq T)$ for
some finite $T>0$, where $\Prob$ is the probability measure under which $X_1,X_2,\dots\sim P$. This is the false alarm probability in $T$ time
units.\xx{do we want to distinguish between $\Prob$ and $P$?}
\item $c_{\ARL}(P;\xi)=\inf\{c>0:\ARL(P;\xi)\geq\gamma\}$ for some
$\gamma>0$. Assuming appropriate continuity, this is the threshold
needed to give an in-control average run length of $\gamma$.
\item $c_{\hit}(P;\xi)=\inf\{c>0:\hit(P;\xi)\leq\beta\}$ for
some $0<\beta<1$. This is the threshold needed
to give a false alarm probability of $\beta$.
\end{itemize}
The latter two quantities are very important in practice, as they are
needed to decide which threshold to use to run a chart. In the
notation we have suppressed the dependence of the quantities on $c$,
$T$, $\gamma$, $\beta$ and $\Delta$.
In the following, $q$ will denote one of $\ARL$, $\hit$, $c_{\ARL}$
or $c_{\hit}$, or simple transformations such as $\log(\ARL)$,
$\logit(\hit)$, $\log(c_{\ARL})$ and $\log(c_{\hit})$, where
$\logit(x)=\log\left(\frac{x}{1-x}\right)$.
The true in-control distribution $P$ and the parameters $\xi=\xi(P)$
needed to run the chart are usually estimated. We assume that we have
past in-control observations $X_{-n},\dots,X_{-1}$ (independent of $X_1,X_2,\dots$), which
we use to estimate the in-control distribution $P$ parametrically or
non-parametrically. We denote this estimate by $\hat P$. The estimate
of $\xi$ will be denoted by $\hat \xi=\xi(\hat P)$. For example, in
the CUSUM control chart of the introduction, $\hat \xi=\hat \mu$ is
the estimated in-control mean.
The observed performance of the chart will depend on the true
in-control distribution $P$ as well as on the estimated parameters
$\hat \xi$ that are used to run the chart. Thus we are interested in
$q(P;\hat\xi)$, the performance of the control chart
\emph{conditional} on $\hat \xi$. This is an unknown quantity as $P$
is not known. Based on the estimator $q(\hat P;\hat\xi)$, we
construct a one-sided confidence interval for this quantity to
guarantee, with high probability, a certain performance for the
chart. We choose to call the interval a confidence interval,
even though the quantity $q(P;\hat\xi)$ is random.
We suggest the following for guaranteeing an upper bound on $q$ (which
is relevant for $q=\hit$, $q=c_{\ARL}$ or $q=c_{\hit}$). For $\alpha\in (0,1)$,
let $p_\alpha$ be a constant such that
$$
\Prob(q(\hat P;\hat\xi) - q(P;\hat\xi) >p_\alpha)=1-\alpha,
$$
assuming that such a $p_{\alpha}$ exists.
Hence,
$$
\Prob(q(P;\hat\xi)< q(\hat P;\hat\xi)-p_{\alpha})= 1-\alpha.
$$
Thus $(-\infty,q(\hat P;\hat\xi)-p_{\alpha})$ could be considered an
exact lower one-sided confidence interval of $q(P;\hat\xi)$.
\xx{Or,
tolerance interval?}
\cc{This is not really a confidence interval in the classical sense -
$q( P;\hat\xi)$ is an
unobserved random quantity... and not just a fixed parameter.
The article Weerahandi (1993, JASA) ``Generalized Confidence Intervals'' might be a useful reference. }
Of course, $p_{\alpha}$ is unknown. We suggest to obtain an
approximation of $p_{\alpha}$ via bootstrapping. In the following,
$\hat P^\ast$ denotes a parametric or non-parametric bootstrap
replicate of the estimated in-control distribution $\hat P$.
We can approximate $p_\alpha$ by $p^\ast_\alpha$ such that
$$\Prob(q(\hat P^\ast;\hat\xi^\ast)-q(\hat P;\hat\xi^\ast)> p^\ast_\alpha|\hat P)=1-\alpha.$$
\cc{alternatively we could use $\hat P(q(\hat P^\ast;\hat\xi^\ast)-q(\hat P;\hat\xi^\ast)> p^\ast_\alpha)=1-\alpha.$
}
Thus
\begin{equation}
\label{eq:onesidedapproxconfint}
(-\infty,q(\hat P;\hat\xi)-p^\ast_\alpha)
\end{equation}
is a one-sided (approximate) confidence interval for $q(P;\hat\xi)$.
In this paper, we will use the following generic algorithm to
implement the bootstrap.
\begin{algorithm}[Bootstrap]
\label{alg:Bootstrap}
\hspace*{2mm}\\[-7mm]
\begin{enumerate}%
\setlength{\itemsep}{0pt}%
\setlength{\parskip}{0pt}%
\item From the past data $X_{-n},\dots,X_{-1}$, estimate
$\hat P$ and $\hat\xi$.
\item Generate bootstrap samples $X^{\ast}_{-n},\dots,X^{\ast}_{-1}$
from $\hat P$. Compute the corresponding estimate $\hat P^{\ast}$
and $\hat \xi^{\ast}$. Repeat $B$ times to get $\hat
P^{\ast}_1,\dots,\hat P^{\ast}_B$ and $\hat \xi^{\ast}_1,\dots,\hat
\xi^{\ast}_B$.
\item Let $p_{\alpha}^{\ast}$ be the $1-\alpha$ empirical quantile of
$q(\hat P^{\ast}_b;\hat\xi^{\ast}_b)-q(\hat P;\hat\xi^{\ast}_b)$, $b=1,\dots,B$.
\end{enumerate}
\end{algorithm}
For guaranteeing a lower bound on $q$, which is for example relevant
for $q=\ARL$, a similar upper one-sided confidence interval can be
constructed.
In a practical situation, the focus would be on deciding which
threshold to use for the control chart to obtain desired in-control
properties. We suggest to use either $q=c_{\ARL}$ or $q=c_{\hit}$, or
log transforms of these, and
then run the chart with the adjusted threshold
\begin{equation}
\label{eq:adjThreshold}
q(\hat P;\hat \xi)-p^{\ast}_{\alpha}.
\end{equation}
This will guarantee that in (approximately) $1-\alpha$ of the
applications of this method, the control chart actually has the
desired in-control properties.
\cc{The following are some comments which are probably not quite relevant to practice.
In some application of control charts the chart parameters $\xi$ are
not estimated but determined according to certain specifications the
process should meet. A typical example would be industrial
applications like monitoring of properties of mass produced units
where there are precise specification of physical properties of the
units which should be monitored. Then $\xi$ may be determined
according to these specifications, and the point of the monitoring is
to detect deviations from the specifications. However, the full
in-control distribution $P$ would usually still be unknown and our
approach would still apply for constructing confidence intervals for
$q(P;\xi_s)$ where $\xi_s$ denotes a specified $\xi$.
Would this in practice be relevant? Or would one also specify $P$,
or at least parts of $P$ like the mean? Could using our approach
here e.g.\ lead to picking a far too large $c$ to get a guaranteed
ARL if $P$ actually is far off from where it ``should be''?
}
\subsection{Specific charts}
\subsubsection{Shewhart charts}
\label{subsubsec:Shewhart}
The one-sided Shewhart chart \citep{Shewhart1931ECo} signals at
$$
\tau=\inf\{t \in \{1,2,\dots\}: f(X_t,\xi)>c\}
$$
for some threshold $c$, where $f$ is some function, $X_t$ is the
observation at time $t$ and $\xi$ are
some parameters.
$X_t$ can be a single
measurement or e.g.\ the average, range or standard deviation of a
specified number of measurements, or some other statistic like a
proportion.
It is common to use a Shewhart chart with a threshold of
the mean plus 3 times the standard deviation, in
this case one would use $c=3$ and $f(x,\xi)=\frac{x-\xi_1}{\xi_2}$
with $\xi_1$ being the mean and $\xi_2$ being the standard deviation.
For two-sided charts one could just use $f(x,\xi)=\frac{|x-\xi_1|}{\xi_2}$.
Conditionally on fixed parameters $ \xi$, the stopping time $\tau$ follows a
geometric distribution with parameter
$p=p(c;P,\xi)=\Prob(f(X_t,\xi)>c)$.
Then the performance measures mentioned in the previous section simplify to
\begin{align*}
\ARL(P;\xi)=&\frac{1}{p(c;P,\xi)},& \hit(P;\xi)=&1-(1-p(c;P,\xi))^T,\\
c_{\ARL}(P;\xi)=&p^{-1}\left( \frac{1}{\gamma}
;P,\xi\right)
\;\;\;\;\;\;
\text{ and}&
c_{\hit}(P;\xi)=&p^{-1}\left( 1-(1-
\beta)^{\frac{1}{T}};P,\xi\right),
\end{align*}
where $p^{-1}(\cdot;P,\xi)$ is the inverse of $p(\cdot;P,\xi)$.
\cc{To get the formula for $c_{\hit}$ set
$\beta=\hit$ in the second item and solve for $c$.}
Suppose that the in-control distribution comes from a parametric
family $P_{\theta}, \theta\in \Theta$. Furthermore, suppose that we
have some way of computing an estimate $\hat\theta$ of $\theta$ based on the
sample.
Then we can use Algorithm \ref{alg:Bootstrap} with $\hat P=P_{\hat \theta}$ to compute a
confidence interval as given by (\ref{eq:onesidedapproxconfint}).
Shewhart charts depend heavily on the tail behaviour of the
distribution of the observations. This is particularly problematic
when the sample size is small and we use non-parametric methods or a
simple non-parametric bootstrap. We thus primarily suggest to use a
parametric bootstrap for Shewhart charts.
\begin{remark}
In certain cases the parametric bootstrap will actually be exact when
$B \to \infty$. This happens when the distribution of
$q(P_{\hat\theta};\hat\xi) - q(P_{\theta};\hat\xi)$ under $P_{\theta}$
does not depend on $\theta$. In particular, this implies that
$q(P_{\hat\theta^\ast};\hat\xi^\ast) - q(P_{\hat\theta};\hat\xi^\ast)$
has the same distribution and $p^\ast_\alpha\to p_\alpha$ as $B\to
\infty$.
As an example, consider the case when
$f(x,\xi)=\frac{x-\xi_1}{\xi_2}$ and $X_t$ follows an
$N(\xi_1,\xi_2^2)$ distribution and $q$ is any of the performance
measures described above. We use $\theta=\xi$ and as estimator $\hat
\xi_1$ we use the sample mean and as estimator $\hat\xi_2$ we use the
sample standard deviation. Then
\begin{align*}
p(c;P_{\!\xi},\hat\xi)
=\Prob_{\!\xi}\!\left(\frac{X_t-\hat\xi_1}{\hat\xi_2}> c\right)
=1-\Phi\left(\frac{c\hat\xi_2+\hat\xi_1-\xi_1}{\xi_2}\right),
\end{align*}
where $\Phi$ is the cdf of the standard normal distribution, and
under $P_\xi$,
$$
\frac{c\hat\xi_2+\hat\xi_1-\xi_1}{\xi_2}=c\frac{\hat\xi_2}{\xi_2}+\frac{\hat\xi_1-\xi_1}{\xi_2}\sim\frac{c}{\sqrt{n-1}}\sqrt{W}+\frac{1}{\sqrt{n}}Z,
$$
where $W\sim\chi_{n-1}^2$ and $Z\sim N(0,1)$ are independent. Thus the
distribution of $p(c;P_\xi,\hat\xi)$, and hence $q(P_{\xi};\hat\xi)$,
is completely known. As $
p(c;P_{\!\hat\xi},\hat\xi)=\Prob_{\!\hat\xi}\left(\frac{X_t-\hat\xi_1}{\hat\xi_2}>
c\right) =1-\Phi(c)$, and thus $q(P_{\!\hat\xi};\hat\xi)$, is not
random, the distribution of $q(P_{\!\hat\xi};\hat\xi) -
q(P_{\xi};\hat\xi)$ also does not depend on any unknown
parameters. Thus the parametric bootstrap is exact in this example.
\end{remark}
\subsubsection{CUSUM charts}
\label{subsubsec:CUSUM}
This section considers the one-sided CUSUM chart \citep{Page1954CIS}.
The classical CUSUM chart was designed to detect a shift of size
$\Delta>0$ in the mean of normally distributed observations. Let $\mu$ and $\sigma$
denote, respectively, the in-control mean and standard
deviation. A CUSUM chart can be defined by
\begin{equation}
\label{eq:discrCUSUM_meanshift}
S_t=\max(0, S_{t-1}+(X_t-\mu - \Delta/2)/\sigma), \quad S_0=0
\end{equation}
with hitting time $\tau=\inf\{t>0: S_t\geq c\}$ for some threshold
$c>0$.
Alternatively, we could drop the scaling and not divide by the
standard deviation $\sigma$ in
(\ref{eq:discrCUSUM_meanshift}). See Chapter 1.4 in
\cite{Hawkins1998csc} for a discussion on scaled versus unscaled
CUSUMs.
More generally, to accommodate observations with general in-control distribution with
density $f_0$ and general out-of-control distribution with density $f_{1}$, it
is optimal in a certain sense \citep{Moustakides1986OST} to modify the
CUSUM chart by replacing $(X_t-\mu - \Delta/2)/\sigma$ by the log
likelihood ratio $\log(f_1(X_t,\theta)/f_0(X_t, \theta))$ such
that the CUSUM chart is
\begin{equation}
\label{eq:discrCUSUM_loglikelihood}
S_t=\max(0, S_{t-1}+\log(f_1(X_t,\theta)/f_0(X_t, \theta))), \quad S_0=0.
\end{equation}
Let $\xi$ denote either $(\mu,\sigma)$ in
(\ref{eq:discrCUSUM_meanshift}) or $\theta$ in
(\ref{eq:discrCUSUM_loglikelihood}). Usually, $\xi$ needs to be
estimated from past data, and we can then use Algorithm
\ref{alg:Bootstrap} to compute a confidence interval
(\ref{eq:onesidedapproxconfint}) for the performance measure
$q(P;\hat\xi)$. For (\ref{eq:discrCUSUM_loglikelihood}) it is most
natural to use a parametric bootstrap with $\hat P=P_{\hat \theta}$,
while for (\ref{eq:discrCUSUM_meanshift}) we can use either a
parametric or a nonparametric bootstrap. In the latter case we let
$\hat P$ be the empirical distribution of $X_{-n},\dots,X_{-1}$, i.e.
in Algorithm \ref{alg:Bootstrap}, $X^{\ast}_{-n},\dots,X^{\ast}_{-1}$
are sampled with replacement from $X_{-n},\dots,X_{-1}$.
\begin{remark}
Similar as for Shewhart charts, this parametric bootstrap is exact
when the distribution of
$q(P_{\hat\theta};\hat\xi)-q(P_{\theta};\hat\xi)$ does not have any
unknown parameters. This is, for instance, the case if we use
(\ref{eq:discrCUSUM_loglikelihood}) for an exponential distribution
with the out-of-control distribution specified as an exponential
distribution with mean $\Delta\lambda$, where $\lambda$ is the
in-control mean. Another example of this is when we have normally
distributed data and use a CUSUM with the increments
$(X_t-\hat\mu)/\hat\sigma-\Delta/2$.
\end{remark}
\section{General theory}
\label{sec:gentheor}
In this section, we show that asymptotically, as the number of past
observations $n$ increases, our procedure works. An established way
of showing asymptotic properties of bootstrap procedures is via a
functional delta method \citep{Vaart1996WCa,Kosorok2008ItE}. Whilst we
will follow a similar route, our problem does not fit directly into
the standard framework, because the quantity of interest, $q(P,\hat
\xi)$, contains the random variable $\hat \xi$.
We present the setup and
the main result in Section \ref{sec:th:main}, followed by examples
(Section \ref{sec:th:examples}).
\cc{The asymptotic development in this section only show that things
do not go badly wrong as $n\to \infty$. They only establish
consistency of the correction/confidence intervals. However, the need to use
these confidence intervals disappears as $n$ increases.}
\subsection{Main theorem}
\label{sec:th:main}
Let $D_q$ be the set in which $P$ and its
estimator $\hat P$ lie, i.e.\ a set describing the potential
probability distribution of our observations. This could be a subset
of $\mathbb{R}^d$ for parametric distributions, the set of cumulative
distribution functions for non-parametric situations, or the set of
joint distributions of covariates and observations. We assume that
$D_q$ is a subset of a complete normed vector space $D$. \cc{Do we
want to /need to assume that $D$ is complete (every Cauchy sequence
converges)? This should not be a problem as $R^k$ and
$l_{\infty}(\mathbb{R})$ are complete metric spaces. } Let $\Xi$ be a
non-empty topological space containing the potential parameters $\xi$
used for running the chart. In our examples, we will let $\Xi\subset
\mathbb{R}^d$ be an open set.
We assume that $\hat P^{\ast}=\hat P^{\ast}(\hat P, W_n)$ is a
bootstrapped version of $\hat P$ based both on the observed data $\hat P$ and
on an independent random vector $W_n$. For example, when resampling
with replacement then $W_n$ is a weight vector of length $n$,
multinomially distributed, that determines how often a given
observation is resampled. In a parametric bootstrap, $W_n$ is the
vector of random variables needed to generate observations from the
estimated parametric distribution.
In the main theorem we will need that the mapping $q:D_q\times
\Xi\to\mathbb{R}$, which returns the property of the chart we are interested
in, satisfies the following extension of Hadamard differentiability.
For the usual definition of Hadamard differentiability see e.g.\
\citep[Section 20.2]{Vaart1998AS}. The extension essentially consists in
requiring Hadamard differentiability in the first component when the second
component is converging.
\begin{definition}
\label{def:haddiffamily}
Let $D,E$ be metric spaces, let $D_f\subset D$ and let $\Xi$ be a
non-empty topological space. \cc{We need at least to be able to speak
about convergence in $\Xi$.} The family of functions
$\{f(\cdot;\xi):D_f \to E: \xi\in \Xi\}$ is called \emph{Hadamard
differentiable at $\theta\in D_f$ around $\xi \in \Xi$ tangentially
to $D_0\subset D$} if there exists a continuous linear map \cc{this is a requirement that also appears in the original definition and which we may be using in our proofs; however, we never prove for our derivatives that they are continuous and linear}
$f'(\theta;\xi):D_0\to E$ such that
$$
\frac{f(\theta+t_nh_n;\xi_n)-f(\theta;\xi_n)}{t_n}\to
f'(\theta;\xi)(h)\quad(n\to\infty)
$$
for all sequences $(\xi_n)\subset \Xi$, $(t_n)\subset \mathbb{R}$, $(h_n)\subset D$
that satisfy $\theta+t_nh_n\in D_f \,\forall n$ and $\xi_n\to \xi$, $t_n\to 0$, $h_n\to h\in D_0$ as $n\to \infty$.
\end{definition}
In the following theorem we understand convergence in distribution, denoted by $\leadsto$, as defined
in \citet[Def 1.3.3]{Vaart1996WCa} or in \citet[p.108]{Kosorok2008ItE}.
\cc{ Let
$(\Omega_n, {\cal A}_n, P_n)$ be a sequence of probability spaces,
let $(\Omega, \cal A, P)$ be a further probability space, let $D$ be
a metric space and let $X_n:\Omega_n\to D$ be a sequence of maps and
let $X:\Omega\to D$ be a Borel measurable map. Then $X_n\leadsto X$
if $\E^{\ast}f(X_n) \to \E f(X)$ for all continuous, bounded $f:D\to
\mathbb{R}$. }
\cc{Outer expectation is defined in \cite{Vaart1996WCa} and
in \cite{Kosorok2008ItE}, essentially $\E^{\ast}X = \inf \{\E Y:
Y\geq X, Y \text{ measurable}\}$.}
\begin{theorem}
\label{th:main}
Let $q:D_q\times \Xi\to \mathbb{R}$ be a mapping, let $P\in D_q$ and let
$\xi:D_q\to \Xi$ be a continuous function.
Suppose that the following conditions are satisfied.
\begin{itemize}%
\setlength{\itemsep}{0pt}%
\setlength{\parskip}{0pt}%
\item[a)] $q$ is Hadamard differentiable at $P$ around $\xi$ tangentially to $D_0$ for some $D_0\subset D$.
\item[b)] $\hat P$ is a sequence of random elements in $D_q$ such that
$
\sqrt{n}(\hat P-P)\leadsto Z
$ as $n\to \infty$
where $Z$ is some tight random element in $D_0$.
\item[c)]
$\sqrt{n}(\hat P^{\ast}-\hat P)\condweakconv{\hat P} Z$ as $n\to\infty$
where $\condweakconv{\hat P}$ denotes weak convergence conditionally on $\hat
P$ in probability as defined in \citet[p.19]{Kosorok2008ItE}. \cc{
i.e. $\sup_{h\in \text{BL}_1}|E_Wh(\hat X_n) - E
h(X)|\stackrel{P}{\to}0$ and $E_Wh(\hat X_n)^{\ast}-E_Wh(\hat
X_n)_{\ast}\stackrel{P}{\to}0$ for all $f\in \text{BL}_1$ where the
subscript $W$ denotes conditional expectation over the weights given
the remaining data. }
\item[d)] The cumulative distribution function of $q'(P;\xi)Z$ is continuous.
\item[e)] Outer-almost surely, the map $W_n\mapsto h(\hat P^{\ast}(\hat P, W_n))$ is measurable for each $n$ and for every continuous bounded function $h:D_q\to \mathbb{R}$.
\item[f)] $q(\hat P; \hat \xi)-q(P;\hat \xi)$ and $p_\alpha^{\ast}$ are random variables, i.e.\ measurable,
where $\hat \xi =\xi(\hat P)$ and $p^{\ast}_{\alpha}=\inf\{t\in \mathbb{R}:
\hat \Prob(q(\hat P^\ast;\hat\xi^\ast)-q(\hat P;\hat\xi^\ast)\leq t)\geq
\alpha\}$.
\end{itemize}
Then
\begin{equation*}
\Prob(q(P;\hat\xi)\in (-\infty, q(\hat P;\hat\xi)-p^{\ast}_{\alpha}))\to 1-\alpha \quad (n\to \infty).
\end{equation*}
\end{theorem}
A similar result holds for upper confidence intervals.
The proof is in Appendix \ref{sec:proof}. The theorem essentially is
an extension of the delta-method. Condition a) ensures the necessary
differentiability. Conditions b) and c) are standard assumptions for
the functional delta method; b) for the ordinary delta method and c)
for the bootstrap version of it. Condition d) ensures that, after
using an extension of the delta-method, the resulting confidence
interval will have the correct asymptotic coverage probability.
Condition e) is a technical measurability condition, which will be
satisfied in our examples. Condition f) is a measurability condition,
which should usually be satisfied.
\subsection{Examples}
\label{sec:th:examples}
The following sections give examples in which Theorem \ref{th:main}
applies. We consider hitting probabilities ($q=\hit$) and thresholds to
obtain certain hitting probabilities ($q=c_{\hit}$).
These examples are
meant to be illustrative rather than exhaustive. For example, other
parametric setups could be considered along similar lines to Section
\ref{sec:cusum-charts-with}. Furthermore, other performance measures such as
$\log(c_{\hit})$ or $\logit(\hit)$ would essentially require application of chain rules
to show differentiability.
\subsubsection{Simple nonparametric setup for CUSUM charts}
\label{sec:theor:ex:CUSUM:nonpar}
We show how the above theorem applies to the CUSUM chart described in
(\ref{eq:discrCUSUM_meanshift}) when using a non-parametric bootstrap
version of Algorithm \ref{alg:Bootstrap}.
Let $D=l_{\infty}(\mathbb{R})$ be the set of bounded functions $\mathbb{R}\to\mathbb{R}$
equipped with the sup-norm $\|x\|=\sup_{t\in \mathbb{R}}|x_t|$. \cc{This is a
Banach space, i.e. a complete normed vector space}
Let $D_q\subset D$ be the set of cumulative distribution functions on
$\mathbb{R}$ with finite second moment.
The parameters needed to run the chart are the mean and the standard deviation of the in-control observations, thus we may choose
$\Xi=\mathbb{R}\times(0,\infty)$ and $\xi:D_q\to \Xi, P\mapsto (\int x P(dx),
\int x^2 P(dx)-(\int x P(dx))^2)$.
As quantities $q$ of interest we are considering hitting probabilities
($q=\hit$) and thresholds ($q=c_{\hit}$) needed to achieve a certain hitting
probability. The probability $\hit:D_q\times \Xi\to \mathbb{R}$ of hitting a
threshold $c>0$ up to step $T>0$ can be written as
$\hit(P;\xi)=\Prob(m(Y) \geq c)$, where
$m(Y)=\max_{i=1,\dots,T}R_i(Y)$ is the maximum value of the chart up
to time $T$,
$R_i(Y)=\sum_{j=1}^iY_j-\min_{0\leq k\leq i}\sum_{j=1}^kY_j$ is the
value of the CUSUM chart at time $i$, $Y=(Y_1,\dots,Y_T)$,
$Y_t=\frac{X_t-\xi_1-\Delta/2}{\xi_2}$ and $X_1,\dots,X_T \sim P$ are the
independent observations. The threshold needed to achieve a certain hitting
probability $\beta \in (0,1)$ is $c_{\hit}: D_q\times \Xi\to \mathbb{R}$,
$c_{\hit}(P;\xi)=\inf\{c>0:\hit(P;\xi)\leq \beta\}$.
The setup for the nonparametric bootstrap is as follows. $W_{n}$ is an
$n$-variate multinomially distributed random vector with probabilities
$1/n$ and $n$ trials. The resampled distribution is $\hat
P^{\ast}=\frac{1}{n}\sum_{j=1}^nW_{nj}\delta_{X_{-j}}$, where $\delta_x$
denotes the Dirac measure at $x$.
The following lemma shows
condition a) of Theorem \ref{th:main},
the Hadamard differentiability of $\hit$ and $c_{\hit}$.
\begin{lemma}
\label{le:HaddiffCUSUMhit}
For every $P\in D_q$, and every $\xi \in \mathbb{R}\times(0,\infty)$, the
function $\hit$ is Hadamard differentiable at $P$ around $\xi$
tangentially to $D_0=\{H:\mathbb{R}\to \mathbb{R}: H\text{ continuous}, \lim_{t\to
\infty}H(t)=\lim_{t\to-\infty}H(t)=0\}$. If, in addition, $P$ has a
continuous bounded positive derivative $f$ with $f(x)\to 0$ as $x\to
\pm \infty$, then $c_{\hit}$ is also Hadamard differentiable at $P$
around $\xi$ tangentially to $D_0$.
\end{lemma}
The proof is in Appendix \ref{sec:haddifhitprobex}, with preparatory results in
Appendix \ref{sec:chain-rule} - \ref{sec:diff-hitt-prob}.
Conditions b) and c) of Theorem \ref{th:main} follow directly from empirical process theory,
see e.g.\ \cite[p.17,Theorems 2.6 and 2.7]{Kosorok2008ItE}.
\cc{To see conditions b) and c) of Theorem \ref{th:main}, we can argue as follows.
In the language of empirical process theory, consider ${\cal
F}=\{\mathbb{R}\to \mathbb{R}, x\mapsto 1_{(-\infty,a]}(x):a\in \mathbb{R}\}$ and let
$l_{\infty}({\cal F})$ be the set of all bounded function ${\cal
F}\to\mathbb{R}$. As $\cal F$ can be identified with $\mathbb{R}$, we can
idenfity $l_{\infty}({\cal F})$ with $l_{\infty}(\mathbb{R})$, the set of
bounded functions $\mathbb{R}\to \mathbb{R}$. By \cite[p.17]{Kosorok2008ItE}, $\cal F$ is
Donsker, i.e.\ if $X_1,\dots,X_n\sim P$ independently, and letting
$P_n=\frac{1}{n}\sum_{i=1}^n\delta_{X_i}$ be the corresponding
empirical measure, then $G_n=\sqrt{n}(P_n-P)\leadsto G$ in
$l_{\infty}(\cal F)$ (or equivalently in $l_{\infty}(\mathbb{R})$) for some
$G$.
thus $G_n$ is considered a random element in $l_{\infty}({\cal
F})$ (or equivalently $l_{\infty}(\mathbb{R})$), via
$\sqrt{n}(P_n-P)(1_{(-\infty,a]})=\sqrt{n}(P_n((-\infty,a])-P((-\infty,a]))$
Now, Theorems 2.6 and 2.7 of \cite{Kosorok2008ItE} give conditional convergence
results for the nonparametric bootstrap, i.e. they show that
$\hat G_n\condweakconv{\hat P} G$ in $l_{\infty}({\cal F})$ and that the sequence $\hat G_n$ is asymptotically measurable.
Sufficient conditions for c) are e.g.\ given in Theorems 3.6.1 and 3.6.2 on p.347 of \cite{Vaart1996WCa}
}
Condition e) is satisfied as well, see bottom of p.189 and after Theorem 10.4 (p.184) of \cite{Kosorok2008ItE}.
Verifying condition d) in full is outside the scope of the present paper.
A starting point could be the fact that by
the Donsker theorem, $Z\sim G\circ P$, where $G$ is a Brownian bridge.
\cc{We would need to consider the derivative in Lemma
\ref{le:diffhitprob} and in Lemma \ref{le:Haddiffinversemap}.}
\subsubsection{CUSUM charts with normally distributed observations}
\label{sec:cusum-charts-with}
In this section, we consider a similar setup to the monitoring based
on (\ref{eq:discrCUSUM_meanshift}) considered in the previous
subsection with the difference that we now use parametric assumptions.
More specifically, the observations $X_i$ follow a normal
distribution with unknown mean $\mu$ and variance $\sigma^2$. We will
use this both for computing the properties of the chart as well as in
the bootstrap, which will be a parametric bootstrap version of
Algorithm \ref{alg:Bootstrap}.
The distribution of the observations can be identified with its
parameters which we estimate by $\hat P = (\hat \mu, \hat \sigma^2)$,
where $\hat \mu=\frac{1}{n}\sum_{i=1}^nX_{-i}$ and $\hat
\sigma^2=\frac{1}{n-1}\sum_{i=1}^n(X_{-i}-\hat \mu)^2$. The set of
potential parameters is $D_q=\mathbb{R}\times(0,\infty)$ which is a subset of
the Euclidean space $D=\mathbb{R}^2$. The parameters needed to run the chart
(\ref{eq:discrCUSUM_meanshift}) are just the same as the one needed to
update the distribution, thus $\Xi=D_q$ and $\xi:D_q\to \Xi,
(\mu,\sigma)\mapsto (\mu,\sigma)$ is just the identity.
As before, we are interested in hitting probabilities within the first
$T$ steps. Using the function $\hit$ defined in the previous
subsection, we can write the hitting probability in this parametric
setup as $\hit^N:D_q\times \Xi\to\mathbb{R}$, $(\mu,\sigma;\xi)\mapsto\hit(
\Phi_{\mu,\sigma^2};\xi)$, where $\Phi_{\mu,\sigma^2}$ is the cdf of
the normal distribution with mean $\mu$ and variance $\sigma^{2}$ and
the superscript $N$ stands for normal distribution. Furthermore,
using $c_{\hit}$ from the previous subsection, the threshold needed to
achieve a given hitting probability is $c_{\hit}^N:D_q\times\Xi\to\mathbb{R}$,
$(\mu,\sigma;\xi)\mapsto c_{\hit}(\Phi_{\mu,\sigma^2};\xi)$.
The resampling is a parametric resampling. To put this in the framework of the main theorem, we let $W_n=(W_{n1},\dots,W_{nn})$, where
$W_{n1},\dots,W_{nn}\sim N(0,1)$ are independent. The
resampled parameters are then $\hat
\mu^{\ast}_n=\frac{1}{n}\sum_{i=1}^nX_{ni}^{\ast}$ and $\hat \sigma^{\ast
2}_n=\frac{1}{n-1}\sum_{i=1}^n(X^{\ast}_{ni}-\hat \mu^{\ast}_n)^2$
where $ X^{\ast}_{ni}=\hat P_2W_{ni}+\hat P_1$.
The following lemma shows that condition a) of Theorem \ref{th:main} is satisfied.
\begin{lemma}
\label{le:HaddiffCUSUMhitNORMAL}
For every $\theta\in \mathbb{R}\times (0,\infty)$ and every $\xi \in \mathbb{R}\times(0,\infty)$,
the functions $\hit^N$ and $c_{\hit}^{N}$ are Hadamard differentiable at $\theta$ around $\xi$.
\end{lemma}
The proof can be found in Appendix \ref{sec:haddifhitprobex}, using again the preparatory results of
Appendix \ref{sec:chain-rule} - \ref{sec:diff-hitt-prob}.
Concerning the other conditions of Theorem \ref{th:main}: Condition b)
can be shown using standard asymptotic theory, e.g.\ maximum likelihood
theory, which will yield that $Z$ is normally distributed. \cc{could
argue via the $(\hat \mu, (n-1)/n\hat \sigma^2)$ being the MLE}
Condition c) is essentially the requirement that the parametric
bootstrap of normally distributed data is working. \cc{This should be
easy to shown by arguing conditionally on the estimators. There may
be something in vdVaart, asymptotic statistics - but he is just
using nonparametric resampling.} As $Z$ is a normally distributed
vector, condition d) holds unless $q'$ equals 0. Condition e) is
satisfied, as the mapping $W_n\mapsto\hat P^{\ast}(\hat P, W_n) =
(\hat\mu_n^{\ast}, \hat\sigma_n^{\ast 2})$ is continuous and hence
measurable.
\subsubsection{Setup for Shewhart charts}
For Shewhart charts, the same setup as in the previous two sections
can be used, the only difference is the choice of $q$. Conditions b),
c) and e) are as in the previous two sections. We conjecture that it is possible to show the Hadamard
differentiability more directly, as the properties are
available in closed form, see Section \ref{subsubsec:Shewhart}.
\cc{
With $G=1-p$,
$\hit(G)=(c\mapsto 1-G(c)^T)$, $\hit'(G)(H)=(c\mapsto -G(c)^{T-1}H(c))$,
$\frac{\partial}{\partial c} \hit(G)(c)=-T G(c)^{T-1}g(c)$
$\ARL(G)=(c\mapsto\frac{1}{1-G(c)})$, $\ARL'(G)(H)=(c\mapsto\frac{1}{(1-G(c))^2}H(c))$ \cc{see \cite[Lemma 3.9.25]{Vaart1996WCa}}
$\frac{\partial}{\partial c}\ARL(G)(c)=\frac{1}{(1-G(c))^2}g(c)$
}
\section{Simulations for homogeneous observations}
\label{sec:simulsingle}
We now illustrate our approach by some simulations using
CUSUM charts. The simulations were done in R \citep{R}.
We use two past sample sizes, $n=50$ and
$n=500$. The in-control distribution of $X_t$ is $N(0,1)$ and
we use 1000 replications and $B=1000$ bootstrap
replications. We employ both the parametric bootstrap and the
nonparametric bootstrap mentioned in the previous sections. For the
parametric bootstrap we used the sample mean and sample standard
deviation of $X_{-n},\dots,X_{-1}$ as estimates for the mean and the standard deviation of
the observations.
For the performance measures $\ARL$, $\log(\ARL)$, $\hit$ and
$\logit(\hit)$ we use a threshold
of $c=3$. For $c_{\ARL}$ we calibrate to an $\ARL$ of $100$
in control and for $c_{\hit}$ we calibrate to a false alarm probability of
$5\%$ in 100 steps.
We use the CUSUM chart (\ref{eq:discrCUSUM_meanshift}) with $\Delta=1$
and $\mu$ and $\sigma$ estimated from the past data. To compute
properties such as $\ARL$ or hitting probabilities, we use a
Markov chain approximation (with 75 grid points), similar to the one
suggested in \cite{BROOK1972atp}\cc{there is a precise description of
a grid that is being used in that paper - I think we are using
something similar but most likely not completely identical}. \cc{We
checked that this gave very good approximations.}
\subsection{Coverage probabilities}
\label{subsec:simulcoverage}
Table \ref{tab:covprob_simnormal} contains coverage probabilities of
nominal 90\% confidence intervals. These are the one-sided lower confidence
intervals given by (\ref{eq:onesidedapproxconfint}), except for
$q=\ARL$ and $\log(ARL)$ where the corresponding upper interval is used.
\begin{table}
\caption{Coverage probabilities of nominal 90\% confidence intervals for CUSUM charts.
\label{tab:covprob_simnormal}}
\begin{center}
\parbox{0.58\textwidth}{
\input{simpaper/tablepaper_stdNormal_CenterScale.tex}\\
The standard deviation of the results is roughly 0.01.
}
\end{center}
\end{table}
In the parametric case, for $n=50$, the coverage probabilities are
somewhat off for untransformed versions, in particular for $q=\ARL$.
Using $\log$ or $\logit$ transformations seems to improve the coverage
probabilities considerably. In the parametric case, for $n=500$, all
coverage probabilities seem to be fine, except for $q=\ARL$, which
although shows some marked improvement compare to $n=50$. In the
nonparametric case, a similar picture emerges, but the coverage
probabilities are a bit worse than in the parametric case.
\begin{remark}
\label{rem:scalingdoesnotmatter}
For $q= \log(c_{\ARL})$ and $q= \log(c_{\hit})$ the division by
$\hat\sigma$ in
(\ref{eq:discrCUSUM_meanshift}) could be skipped without making a
difference to the coverage probabilities. Indeed, the division by
$\hat\sigma$ just scales the chart (and the resulting threshold) by
a multiplicative factor, which is turned into an additive factor by
$\log$ and which then cancels out in our adjustment.
\end{remark}
\subsection{The benefit of an adjusted threshold}
In this section, we consider both the in- and out-of-control
performance of CUSUM charts when adjusting the threshold $c$ to give a
guaranteed in-control $\ARL$ of 100. Setting the threshold is, in our
opinion, the most important practical application of our method.
\begin{figure}[tb]
\centering
\includegraphics[width=\linewidth]{simpaper/adjusted_unadjusted_ARL_CenterScale_boxplot.pdf}
\caption{Distribution of the conditional $\ARL$ for CUSUMs in a
normal distribution setup. Thresholds are calibrated to an
in-control $\ARL$ of 100. The adjusted thresholds have a
guarantee of 90\%. A log transform is used in the calibration.
The boxplots show the 2.5\%, 10\%, 25\%, 50\%, 75\%, 90\% and
97.5\% quantiles. The white boxplots are in-control, the gray
boxplots out-of-control.}
\label{fig:adjusted_unadjusted_ARL}
\end{figure}
Figure \ref{fig:adjusted_unadjusted_ARL} shows average run lengths
for both the unadjusted threshold $c(\hat P;\hat \mu, \hat \sigma)$ and the
adjusted threshold $\exp(\log( c(\hat P;\hat \mu, \hat \sigma))-p^{\ast}_{0.1})$, where $p^{\ast}_{0.1}$ is computed via the parametric
bootstrap using $q=\log(c_{\ARL})$. Thus, with 90\% probability, the adjusted threshold should
lead to an $\ARL$ that is above 100. In this and in all following
simulations,
the out-of-control ARL refers
to the situation where the chart is out-of-control from the
beginning, i.e.\ from
time 0 onwards.
For the unadjusted threshold, the desired in-control average run
length is only reached in roughly half the cases. More importantly,
for $n=50$, the probability of having an in-control $\ARL$ of below $50$
is greater than 20\%.
With the adjusted threshold we should get an average run length of at
least 100 in 90\% of the cases. This is achieved. The
out-of-control $\ARL$ using the adjusted thresholds increases only
slightly compared to the unadjusted version.
Similarly to Remark \ref{rem:scalingdoesnotmatter}, removing the
scaling by $\hat \sigma$ in (\ref{eq:discrCUSUM_meanshift}) would not
change the results of this section.
\subsection{Nonparametric bootstrap - advantages and disadvantages}
In this section, we compare the parametric and the non-parametric
bootstrap. We consider CUSUM charts that are calibrated to an
in-control average run length of 100 assuming a normal distribution.
We use the adjusted threshold $\exp(\log( c_{\ARL}(\hat P;\hat
\mu, \hat \sigma))-p^{\ast}_{0.1})$.
Figure \ref{fig:par_nonpar_ARL} shows the distribution of $\ARL$ for
$n=50$ and $n=500$ for both the parametric bootstrap that assumes a normal
distribution of the updates and the nonparametric bootstrap. We consider both a
correctly specified model where $X_t\sim N(0,1)$ as well as two
misspecified models where $X_t\sim \text{Exponential}(1)$ and $\sqrt{20}X_t\sim
\chi^2_{10}$ (all of the $X_t$ have variance 1). We show both the in- as well as the
out-of-control performance of the charts.
\begin{figure}[tb]
\centering
\includegraphics[width=\linewidth]{simpaper/par_nonpar_ARL_CenterScale_boxplot.pdf}
\caption{Effects of misspecification. Thresholds are calibrated to
an in-control ARL of 100 and adjusted to the estimation error with
a guarantee of 90\%. A log transform is used in the
calibration. The white boxplots are in-control, the gray
boxplots are out-of-control. The boxplots show
the 2.5\%, 10\%, 25\%, 50\%, 75\%, 90\% and 97.5\% quantiles.}
\label{fig:par_nonpar_ARL}
\end{figure}
In the correctly specified model ($X_t\sim N(0,1)$), the performance
of the parametric and the non-parametric chart seems to be almost
identical. The only difference is a slightly worse in-control
performance for the non-parametric chart for $n=50$.
In the misspecified model with $X_t\sim \text{Exponential}(1)$, the
parametric chart does not have the desired in-control
probabilities. The non-parametric chart seems to be doing well, in
particular for $n=500$. We have a similar results in the other
misspecified model, with $\sqrt{20}X_t\sim \chi^2_{10}$.
\section{Regression models}
\label{sec:regmod}
In many monitoring situations, the units being monitored are heterogeneous,
for instance when monitoring patients at hospitals or bank customers.
To make sensible monitoring systems in such situations, the explainable
part of the heterogeneity should be accounted for by relevant
regression models. The resulting charts are often called risk
adjusted, and an overview of some such charts can be found in \cite{Grigg2004oor}.
To run risk adjusted charts, the regression model needs to be estimated
based on past data, and this estimation needs to be accounted for. Our
approach for setting up charts with a guaranteed performance applies
also to risk adjusted charts,
and we will in particular look at linear, logistic and survival
models.
\subsection{Linear models}
\label{subsec:linmod}
Suppose we have independent observations $(Y_1,X_1),$ $(Y_2,X_2)$, $\ldots$,
where $Y_i$ is a response of interest and $X_i$ is a corresponding
vector of covariates, with the first component usually equal to 1.
Let $P$ denote the joint distribution of $(Y_i,X_i)$ and suppose that
in control $\E(Y_i|X_i)=X_i\xi$. From some observation
$\kappa$ there is a shift in the mean response to
$\E(Y_i|X_i)=\Delta+X_i\xi$ for $i=\kappa,\kappa+1,\dots$.
Monitoring schemes for detecting changes in regression models can
naturally be based on residuals of the model, see for instance
\cite{Brown1975TfT} and \cite{Horvath2004Mci}.
We can, for instance,
define a CUSUM to monitor changes in the conditional mean of $Y$ by
$$S_t=\max(0, S_{t-1}+Y_t-X_t \xi - \Delta/2), \quad S_0=0, $$
with hitting time $\tau=\inf\{t>0: S_t\geq c\}$ for some threshold
$c>0$.
In a similiar manner we could also set up charts for
monitoring changes in other components of $\xi$.
The parameter vector $\xi$ is estimated from past in
control data, e.g.\ by the standard least squares estimator. We
suggest to use a nonparametric version of the general Algorithm
\ref{alg:Bootstrap} with $\hat P$ being the
empirical distribution putting weight $1/n$ on each of the past
observations $(Y_{-n},X_{-n}),\dots,(Y_{-1},X_{-1})$. Resampling is
then equivalent to resampling
$(Y^{\ast}_{-n},X^{\ast}_{-n}),\dots,(Y^{\ast}_{-1},X^{\ast}_{-1})$ by
drawing with replacement from $\hat P$.
The suggested method should work even if the linear model is misspecified,
i.e.\ $\E(Y_i|X_i)=X_i\xi$ does not necessarily hold. The nonparametric
bootstrap should take this into account.
An analogous approach can be used for Shewhart charts. In settings
where it is reasonable to consider the covariate vector to be
non-random one could alternatively use bootstrapping of residuals, see
for example \cite{Freedman1981BRM}.
\subsubsection{Theoretical considerations}
Obtaining precise results is more demanding than in the examples
without covariates in Section~\ref{sec:th:examples}. We only
give an idea of the setup that might be used.
The set of distributions of the observations $D_q$ can be chosen as the
set of cdfs on $\mathbb{R}^{d+1}$ with finite second moments, where $d$ is the dimension of the covariate. The first cdf corresponds to the responses, the others to the covariates. $D_q$ is contained
in the vector space $D=l_{\infty}(\mathbb{R}^{d+1})$, the set of bounded functions $\mathbb{R}^{d+1}\to \mathbb{R}$.
The parameters needed to run the chart are the regression coefficients contained in the set $\Xi=\mathbb{R}^d$.
These parameters are obtained from the distribution of the observations via $\xi:D_q\to \Xi$, $F\mapsto
(E(X^TX))^{-1}E(X Y)$ where $(Y,X)\sim F$ where $X$ is considered to
be a row vector.
We conjecture that the conditions of Theorem \ref{th:main} are broadly
satisfied if the cdf of $Y-X\xi$ is differentiable and if for
the property $q$ we use hitting probabilities or thresholds to
achieve a given hitting probability. In particular, it should be
possible to show Hadamard differentiability similarly to Lemma
\ref{le:HaddiffCUSUMhit}: write $q$ as concatenation of two functions
and use the chain rule in Lemma \ref{le:chainrule}. The first mapping
returns the distribution of the updates of the chart depending on
$F\in D_q$ and $\xi\in \Xi$ via $(F;\xi)\mapsto {\cal L}(Y-X\xi-
\Delta)$, where ${\cal L}$ denotes the law of a random variable. The
second takes the distribution of the updates and returns the property of
interests. The differentiability of the second map has been shown in
Lemmas \ref{le:Haddiffinversemap} and \ref{le:diffhitprob}.
\subsubsection{Simulations}
\label{example:CUSUMLinReg}
We illustrate the performance of the bootstrapping scheme using a
CUSUM and the linear in-control model
$Y=X_{1}+X_{2}+X_3+\epsilon$. Let $\epsilon\sim N(0,1)$, $X_{1}\sim
\text{Bernoulli}(0.4)$, $X_2\sim U(0,1)$ and $X_3\sim N(0,1)$, where
$X_1,X_2,X_3$ and $\epsilon$ are all independent. The out-of-control
model is $Y=1+X_{1}+X_{2}+X_3+\epsilon$,
i.e. $\Delta=1$. Figure~\ref{fig:regression_ARL} shows the distribution
of the attained ARL for CUSUMs with thresholds calibrated to give an
in control ARL of 100. We see that the behaviour of the adjusted
versus unadjusted thresholds are very similar to what we observed for
the simpler model in Figure~\ref{fig:adjusted_unadjusted_ARL}. The
coverage probabilities obtained for this regression model, not
reported here, are also very similar to the covarage probabilities
reported in Table~\ref{tab:covprob_simnormal}, though with a tendency
to be slightly worse.
\begin{figure}[tb]
\centering
\includegraphics[width=\linewidth]{simpaper/regression_ARL_boxplot.pdf}
\caption{Distribution of the conditional $\ARL$ for CUSUMs in a
linear regression setup. Thresholds are calibrated to an
in-control $\ARL$ of 100. A log transform is used in the
calibration. The adjusted thresholds have a guarantee of
90\%. The white boxplots are in control, the gray
out-of-control. The boxplots show the 2.5\%, 10\%, 25\%, 50\%,
75\%, 90\% and 97.5\% quantiles.}
\label{fig:regression_ARL}
\end{figure}
\subsection{Logistic regression}
\label{subsec:logreg}
Control charts, in particular CUSUM charts, based on logistic
regression models are popular for modelling of binary
outcomes in medical contexts. See e.g.\ \cite{Lie1993nsp}, \cite{Steiner2000Msp},
\cite{Grigg2004oor} and \cite{Woodall2006Tuo}.
Suppose we have independent observations $(Y_1,X_1),(Y_2,X_2),\ldots,$
where $Y_i$ is a binary response variable and $X_i$ is a corresponding
vector of covariates. Further, suppose that in control the log odds
ratio is $\logit(\Prob(Y_i=1|X_i))=X_i\xi$, and that from some observation
$\kappa$ there is a shift in the log odds ratio to
$\logit(\Prob(Y_i=1|X_i))=\Delta+X_i\xi$ for $i=\kappa,\kappa+1,\dots$
A CUSUM to monitor changes in the odds ratio can be defined by \citep{Steiner2000Msp}
$$S_t=\max(0, S_{t-1}+R_t), \quad S_0=0, $$
where $R_t$ is the log likelihood ratio between the in-control and out-of-control model for observation $t$. More precisely
$$
\exp(R_t)=\frac{\exp(\Delta+X_t\xi)^{Y_t}/(1+\exp(\Delta+X_t\xi))}{\exp(X_t\xi)^{Y_t}/(1+\exp(X_t\xi))}
=\exp(Y_t\Delta)\frac{1+\exp(X_t\xi)}{1+\exp(\Delta+X_t\xi)}.
$$
The parameter vector $\xi$ is estimated from past in-control
data by e.g.\ the standard maximum likelihood estimator. The same
nonparametric bootstrap approach as described for the linear model in
Section~\ref{subsec:linmod} can now be applied to this CUSUM based on
this logistic regression model. Moreover, this approach would also
apply to control charts based on other generalized linear models, for
instance Poisson regression models for monitoring count data. The only
amendment needed is to replace $R_t$ by the relevant log likelihood
ratio.
We have run simulations, not reported here, based on the same covariate
specifications as in Section~\ref{example:CUSUMLinReg}. The results
are similar to the results for the linear model of
Section~\ref{example:CUSUMLinReg}.
\subsection{Survival analysis models}
Recently, risk adjusted control charts based on survival models have
started to appear, see
\cite{Biswas2008rCi,Sego2009Rmo,Steiner2009ras,Gandy2010ram}. In none
of these papers any adjustment for estimation error is done, but
\cite{Sego2009Rmo} are illustrating, by simulations, the impact of
estimation error on the attained average run length for the
accelerated failure time model based CUSUM studied in their paper.
In the following, we provide a brief simulation example of our
adjustment in a survival setup where we use the methods described in
\cite{Gandy2010ram}.
We observe the survival of individuals over a fixed time interval of
length $n$ (we will use $n=100$ and $n=500$). Individuals arrive at
times $B_i$ (in our simulation according to a Poisson process with
rate $1$), and survive for $T_i$ time units. Individuals may arrive
before the observation interval, as long as $B_i+T_i$ is after the
start of the observation interval. Right-censoring, at $C_i$ time
units after arrival, is taking place after a maximum follow-up time of
$t=60$ time units or after the individuals leave the observation
interval. In the simulation, the true hazard rate of $T_{i}$ is
$h_i(t)= 0.1\exp( X_{1i}+X_{2i})$, where $X_{1i}\sim
\text{Bernoulli}(0.4)$ and $X_{2i}\sim N(0,1)$ are covariates.
Based on the observed data we fit a Cox proportional hazard model
with $X_{1i}$ and $X_{2i}$ as covariates and nonparametric baseline,
giving estimates $\hat \beta$ for the covariate effects and $\hat
\Lambda_0(t)$ for the the integrated baseline.
We use the CUSUM chart described in
\cite{Gandy2010ram} against a proportional alternative with
$\rho=1.25$. The parameters
needed to run the chart are $\xi=(\beta, \Lambda_0)$ estimated by
$\hat\xi=(\hat \beta, \hat \Lambda_0)$.
To be precise, the chart signals at time
$\tau=\inf\{t>0:S(t)\geq c\}$, where
$S(t)=R(t)-\inf_{s\leq t}R(s)$,
$
R (t ) = \log(\rho ) N (t ) - (\rho - 1)
\Lambda(t),
$
$N (t )$ is the number of events until time $t$ and $\Lambda(t ) =
\sum_{i} \exp( \beta_1X_{i1}+ \beta_2X_{2i})
\Lambda_0(\min((t-B_i)^{+},T_i,C_i))$.
We are interested in finding a threshold that gives a desired
hitting probability, i.e. we use $q=c_{\hit}$. We compute
$c_{\hit}(P,\xi)$ via simulations (simulate new data from $P$ and run
the chart with $\xi$). We estimate the threshold needed to get a 10\%
false alarm probability in $n$ time units in control, by the 90\% quantile of 500
simulations of the maximum of the chart.
To resample, we resample individuals with replacement. We use 500
bootstrap samples. Figure \ref{fig:hitprobsurvanal} shows the
distribution of the resulting hitting probabilities based on 500
simulated observation intervals.
\begin{figure}
\centering
\includegraphics[width=\linewidth]{simpaper/adjusted_unadjusted_coxhitprob_boxplot.pdf}
\caption{Distribution of the conditional hitting probability for
survival analysis CUSUMs. Thresholds are calibrated to an in-control
hitting probability of 0.1. The adjusted thresholds have a
guarantee of 90\%. The white boxplots are in control, the gray
out-of-control. The boxplots show the 2.5\%, 10\%, 25\%, 50\%, 75\%,
90\% and 97.5\% quantiles. \label{fig:hitprobsurvanal}}
\end{figure}
In control, without the adjustment, the desired false
alarm probability of 0.1 is only reached in roughly 60\% of the cases. The
bootstrap correction seems to work fine, leading to a false alarm probability
of at most 10\% in roughly 90\% of the cases. As expected, increasing the
length of the fitting period and the length of time the chart is run
from $n=100$ to $n=500$ results in higher out-of-control hitting probabilities.
If the length of the fitting period and the deployment period of the chart differ then
a somewhat more complicated resampling procedure needs to be used.
For example, one could
resample arrival times and survival times/covariates separately.
The former could be done by assuming a Poisson process as arrival time and the
latter either by resampling with replacement or by sampling from an estimated
Cox model and an estimated censoring distribution.
\cc{
In the survival analysis case with a proportional alternative, the chart is based on
$$
R(t) = \log(\rho ) N (t )- (\rho - 1)\hat \Lambda (t ),
$$
After the time transformation of $N$ to the standard Poisson process $\tilde N$ this becomes
\begin{align*}
\tilde R(t) = R(\Lambda^{-1}(t))=\log(\rho)\tilde
N(t)-(\rho-1)\hat\Lambda(\Lambda^{-1}(t))
\end{align*}
Thus the nice Markov-approximation will not work $\hat
\Lambda$ and $\Lambda$ will not have independent
increments. Therefore we needed to simulate.
}
\section{Conclusions and discussion}
\label{sec:conclusion}
We have presented a general approach for handling estimation error in
control charts with estimated parameters and unknown in-control
distributions. Our suggestion is, by bootstrap methods, to tune the
monitoring scheme to guarantee, with high probability, a certain
conditional in-control performance (conditional on the estimated
in-control distribution). If we apply a nonparametric bootstrap, the
approach is robust against model specification error.
In our opinion, focusing on a guaranteed conditional in-control performance is
generally more relevant than focusing on some average
performance, as an estimated chart usually is run for some time without
independent reestimation. Our approach can also easily be adapted to
make for instance bias adjustments. Bias adjustments, in
contrast to guaranteed performance, tend to
be substantially influenced by tail behaviour for heavy
tailed distributions which for instance the average run length has.
This implies that the bias adjustments need not be useful in the
majority of cases as the main effect of the adjustment is to adjust
the tail behaviour.
We have in particular demonstrated our approach for various variants
of Shewhart and CUSUM charts, but the general approach will
apply to other charts as well. The method is generally
relevant when the in-control distribution is unknown and the conditions of
Theorem~\ref{th:main} hold. We conjecture that this will be the case
for many of the most commonly used control charts.
\cc{for instance be the case for charts like EWMA charts \citep{Roberts1959CCT}, general
likelihood ratio based charts \cite{Frisen1991Ops,Frisen2003SSO}, the
Sets method \citep{Chen1978SSC,Grigg2004ARA}}
Numerous extensions of control charts to other settings exist, for example
to other regression
models, to autocorrelated data, to multivariate data.
We do conjecture that our approach will also apply in
many of these settings.
\small
|
\subsection{Recipe Pair Prediction}
The goal of our prediction task is: \textit{given a pair of similar recipes, determine
which one has higher average rating than the other}. This task is
designed particularly to help users with a specific dish or meal in mind, and who
are trying to decide between several recipe options for that dish.
\textbf{Recipe pair data.} The data for this prediction task consists of pairs of
similar recipes. The reason for selecting similar recipes, with high ingredient overlap, is that while apples may be quite comparable to oranges in the context of recipes, especially if one is evaluating salads or desserts, lasagna may not be comparable to a mixed drink. To derive pairs of related recipes, we computed similarity with a cosine similarity between the ingredient
lists for the two recipes, weighted by the inverse document frequency,
$log(\# \:of \:recipes / \# \:of \:recipes \:containing \:the \:ingredient)$. We
considered only those pairs of recipes whose cosine similarity
exceeded 0.2.
The weighting is intended to identify higher similarity among recipes sharing more distinguishing ingredients, such as Brussels sprouts, as opposed to recipes sharing very common ones, such as butter.
A further challenge to obtaining reliable relative rankings of recipes is variance introduced by having different users choose to rate different recipes. In addition, some users might not have a sufficient number of reviews under their belt to have calibrated their own rating scheme. To control for variation introduced by users, we examined recipe pairs where the {\em same} users are rating both recipes and are collectively expressing a preference for one recipe over another. Specifically, we generated 62,031 recipe pairs ($a,b$) where $rating_i(a)$ > $rating_i(b)$, for at least 10 users $i$, and over 50\% of users who rated both recipe $a$ and recipe $b$. Furthermore, each user $i$ should be an active enough reviewer to have rated at least 8 other recipes.
\textbf{Features.} In the prediction dataset, each observation consists of a set of predictor
variables or features that represent information about two recipes, and the
response variable is a binary indicator of which gets the higher rating on
average. To study the key aspects of recipe information, we constructed
different set of features, including:
\begin{itemize}
\vspace{-.6em}
\item Baseline: This includes cooking methods, such as chopping, marinating, or grilling, and
cooking effort descriptors, such as preparation time in minutes, as well as the number of
servings produced, etc. These features are considered as primary information about
a recipe and will be included in all other
feature sets described below.
\vspace{-.6em}
\item Full ingredients: We selected up to 1000 popular ingredients to
build a ``full ingredient list''. In this feature set, each observed
recipe pair contains a vector with entries indicating whether an ingredient
from the full list is present in either recipe in the pair.
\vspace{-.6em}
\item Nutrition: This feature set does not include any ingredients but
only nutrition information such the total caloric content, as well
as quantities of fats, carbohydrates, etc.
\vspace{-.6em}
\item Ingredient networks: In this set, we replaced the full ingredient
list by structural information extracted from different ingredient networks, as
described in Sections~\ref{sec:complementnet} and~\ref{sec:substitutenet}. Co-occurrence is treated separately as a raw count, and a complementarity, captured by the PMI.
\vspace{-.6em}
\item Combined set: Finally, a combined feature set is constructed to
test the performance of a combination of features, including baseline, nutrition and ingredient networks.
\end{itemize}
\vspace{-.6em}
To build the ingredient network feature set, we extracted the following two types of structural information from
the co-occurrence and substitution networks, as well as the complement
network derived from the co-occurrence information:
\textit{Network positions} are calculated to represent how a recipe's
ingredients occupy positions within the networks. Such position
measures are likely to inform if a recipe contains any ``popular'' or
``unusual'' ingredients. To calculate the position measures, we first
calculated various network centrality measures, including degree
centrality, betweenness centrality, etc., from the ingredient
networks. A centrality measure can be represented as a vector
$\vec{g}$ where each entry indicates the centrality of an
ingredient. The network position of a recipe, with its full ingredient list represented as a
binary vector $\vec{f}$, can be summarized by
$\vec{g}^{T}\cdot\vec{f}$, i.e., an aggregated centrality measure
based on the centrality of its ingredients.
\textit{Network communities} provide information about which ingredient is
more likely to co-occur with a group of other ingredients in the
network. A recipe consisting of ingredients that are frequently used
with, complemented by or substituted by certain groups may be predictive of the ratings the recipe will receive. To obtain the network community information,
we applied latent semantic analysis (LSA) on recipes.
We first factorized each ingredient network, represented by matrix
$W$,
using singular value decomposition
(SVD). In the matrix $W$, each entry $W_{ij}$ indicates whether ingredient
$i$ co-occurrs, complements or substitues ingredient $j$.
Suppose
$W_k= U_k\Sigma_kV_k^T$ is a rank-$k$ approximation of $W$, we can
then transform each recipe's full ingredient list using the
low-dimensional representation, $\Sigma_k^{-1} V_k^T \vec{f}$, as
community information within a network. These
low-dimensional vectors, together with the vectors of network
positions, constitute the ingredient network features.
\begin{figure}[t!]\centering
\includegraphics[trim=0 30 0 0,width=1\columnwidth]{figs/predres.pdf}
\caption{\label{fig:predres} Prediction performance. The nutrition
information and ingredient networks are more effective features than
full ingredients. The ingredient network features lead to impressive performance, close to
the best performance. }
\vspace{-1em}
\end{figure}
\begin{figure}[t!]\centering
\includegraphics[trim=0 30 0 0,width=1\columnwidth]{figs/relimp.pdf}
\caption{\label{fig:relimp} Relative importance of features in the
combined set. The individual items from nutrition information are very
indicative in differentiating highly rated recipes, while most of the
prediction power comes from ingredient networks.}
\vspace{-1em}
\end{figure}
\textbf{Learning method.} We applied discriminative machine
learning methods such as support vector machines (SVM)~\cite{cortes1995support} and stochastic
gradient boosting trees~\cite{friedman2002stochastic} to our prediction
problem. Here we report and discuss the detailed results based on the gradient boosting tree model.
Like SVM, the gradient boosting tree model seeks a parameterized
classifier, but unlike SVM that considers all the
features at one time, the boosting tree model considers a set
of features at a time and iteratively combines them according to their
empirical errors. In practice, it not only has competitive performance
comparable to SVM, but can serve as a feature ranking procedure~\cite{lu2006coupling}.
In this work, we fitted a stochastic gradient boosting tree model with 8
terminal nodes under an exponential loss function. The dataset is
roughly balanced in terms of which recipe is the higher-rated one
within a pair. We randomly divided the dataset into a training set
(2/3) and a testing set (1/3).
The prediction performance is evaluated
based on accuracy, and the feature performance is evaluated in terms
of relative importance~\cite{hastie2005elements}. For each single
decision tree, one of the input variables, $x^j$, is used to partition
the region associated with that node into two subregions in order to
fit to the response values.
The squared relative importance of variable $x^j$ is the sum of such squared improvements over all internal nodes for which it was chosen as the splitting variable, as:
\[
imp(j)=\sum_{k} \hat{i}_k^2 I(\mbox{splits on } x^j)
\]
where $\hat{i}_k^2$ is the empirical improvement by the $k$-th node splitting on
$x^j$ at that point.
\subsection{Results}
\begin{figure}[!]\centering
\includegraphics[trim=0 30 0 0,width=1\columnwidth]{figs/relimp_network.pdf}
\caption{\label{fig:relimpnet} Relative importance of features
representing the network structure. The substitution network has the strongest contribution
($39.8\%$) to the total importance of network features, and it also has more influential features in
the top 100 list, which suggests that the substitution
network is complementary to other features.}
\vspace{-1em}
\end{figure}
\begin{figure}[t!]\centering
\includegraphics[trim=0 30 0 0,width=1\columnwidth]{figs/relimp_nutrition.pdf}
\caption{\label{fig:relimpnut} Relative importance of features from
nutrition information. The carbs item is the
most influential feature in predicting higher-rated recipes.}
\vspace{-1em}
\end{figure}
The overall prediction performance is shown in
Fig.~\ref{fig:predres}. Surprisingly, even with a full list of ingredients, the
prediction accuracy is only improved from .712 (baseline) to .746. In contrast,
the nutrition information and ingredient networks are more effective
(with accuracy .753 and .786, respectively). Both of them have much lower
dimensions (from tens to several hundreds), compared with the full
ingredients that are represented by more than 2000 dimensions (1000 ingredients per recipe in the pair). The
ingredient network features lead to impressive performance, close to
the best performance given by the combined set (.792), indicating the power of network structures
in recipe recommendation.
Figure~\ref{fig:relimp} shows the influence of different
features in the combined feature set. Up to 100 features with
the highest relative importance are shown. The importance of a feature group is
summarized by how much the total importance is contributed by all
features in the set. For example, the baseline consisting of cooking
effort and cooking methods contribute $8.9\%$ to the overall
performance. The individual items from nutrition information are very
indicative in differentiating highly-rated recipes, while most of the
prediction power comes from ingredient networks ($84\%$).
Figure~\ref{fig:relimpnet} shows the top 100 features from the three
networks. In terms of the total importance of ingredient network
features, the substitution network has slightly stronger contribution
($39.8\%$) than the other two networks, and it also has more
influential features in the top 100 list. This suggests that
the structural information extracted from the substitution network is
not only important but also complementary to information from other aspects.
\enlargethispage{\baselineskip}
Looking into the nutrition information (Fig.~\ref{fig:relimpnut}), we found that carbohydrates are the
most influential feature in predicting higher-rated recipes. Since
carbohydrates comprise around $50\%$ or more of total calories, the
high importance of this feature interestingly suggests that a recipe's
rating can be influenced by users' concerns about nutrition and diet. Another
interesting observation is that, while individual nutrition items are
powerful predictors, a higher prediction accuracy can be reached by
using ingredient networks alone, as shown in
Fig.~\ref{fig:predres}. This implies the information about nutrition may
have been encoded in the ingredient network structure, e.g. substitutions of less healthful ingredients with ``healthier'' alternatives.
\begin{figure}[t!]\centering
\includegraphics[trim=0 30 0 0,width=1\columnwidth]{figs/pred_reduced_dimension.pdf}
\caption{\label{fig:dimension} Prediction performance over reduced
dimensionality. The best performance is given by
reduced dimension $k=50$ when combining all three networks. In
addition, using the information about the complement network alone is more
effective in prediction than using other two networks.}
\vspace{-1em}
\end{figure}
\begin{figure}[t!]\centering
\includegraphics[clip=true,trim=30 114 10 20,width=1\columnwidth]{figs/svn_heatmap_subs.pdf}
\caption{\label{fig:svdsubs} Influential substitution communities. The matrix shows the most influential feature
dimensions extracted from the substitution network. For each
dimension, the six representative ingredients with the highest intensity values
are shown, with colors indicating their
intensity. These features suggest that the communities of ingredient substitutes, such
as the sweet and oil in the first dimension, are particularly
informative in prediction. }
\vspace{-1em}
\end{figure}
Constructing the ingredient network feature involves reducing
high-dimensional network information through SVD, as described in the previous
section. The dimensionality can be determined by cross-validation. As
shown in Fig.~\ref{fig:dimension}, features with a very large
dimension tend to overfit the training data. Hence we chose
$k=50$ for the reduced dimension of all three networks. The figure also
shows that using the information about the complement network alone is more
effective in prediction than using either the co-occurrence and substitute networks, even in the
case of low dimensions. Consistently, as shown in terms of relative
importance (Fig.~\ref{fig:relimpnet}),
the substitution network alone is not the most effective, but it provides
more complementary information in the combined feature set.
In Figure~\ref{fig:svdsubs} we show the most representative ingredients
in the decomposed matrix derived from the substitution network. We
display the top five influential dimensions, evaluated based on the
relative importance, from the SVD resultant matrix $V_k$, and in each
of these dimensions we extracted six representative ingredients based on
their intensities in the dimension (the squared entry values). These
representative ingredients suggest that the communities of ingredient
substitutes, such as the sweet and oil substitutes in the first
dimension or the milk substitutes in the second dimesion (which is
similar to the cluster shown in Fig.~\ref{fig:subclusters}), are particularly
informative in predicting recipe ratings.
To summarize our observations, we find we are
able to effectively predict users' preference for a recipe, but the
prediction is not through using a full list of ingredients. Instead,
by using the structural information extracted from the relationships
among ingredients, we can better uncover users' preference about
recipes.
\section{Introduction}
The web enables individuals to collaboratively share knowledge and recipe websites are one of the earliest examples of collaborative knowledge sharing on the web.
Allrecipes.com, the subject of our present study, was founded in 1997, years ahead of other collaborative websites such as the Wikipedia.
Recipe sites thrive because individuals are eager to share their recipes, from family recipes that had been passed down for generations, to new concoctions that they created that afternoon, having been motivated in part by the ability to share the result online. Once shared, the recipes are implemented and evaluated by other users, who supply ratings and comments.
The desire to look up recipes online may at first appear odd given that tombs of printed recipes can be found in almost every kitchen. The Joy of Cooking~\cite{rombauer1997joy} alone contains 4,500 recipes spread over 1,000 pages. There is, however, substantial additional value in online recipes, beyond their accessibility. While the Joy of Cooking contains a single recipe for Swedish meatballs, Allrecipes.com hosts ``Swedish Meatballs I'', ``II'', and ``III'', submitted by different users, along with 4 other variants, including ``The Amazing Swedish Meatball''. Each variant has been reviewed, from 329 reviews for ``Swedish Meatballs I" to 5 reviews for ``Swedish Meatballs III". The reviews not only provide a crowd-sourced ranking of the different recipes, but also many suggestions on how to modify them, e.g. using ground turkey instead of beef, skipping the ``cream of wheat'' because it is rarely on hand, etc.
The wealth of information captured by online collaborative recipe sharing sites is revealing not only of the fundamentals of cooking, but also of user preferences. The co-occurrence of ingredients in tens of thousands of recipes provides information about which ingredients go well together, and when a pairing is unusual. Users' reviews provide clues as to the flexibility of a recipe, and the ingredients within it. Can the amount of cinnamon be doubled? Can the nutmeg be omitted? If one is lacking a certain ingredient, can a substitute be found among supplies at hand without a trip to the grocery store? Unlike cookbooks, which will contain vetted but perhaps not the best variants for some individuals' tastes, ratings assigned to user-submitted recipes allow for the evaluation of what works and what does not.
In this paper, we seek to distill the collective knowledge and preference about cooking through mining a popular recipe-sharing website. To extract such information, we first parse the unstructured text of the recipes and the accompanying user reviews. We construct two types of networks that reflect different relationships between ingredients, in order to capture users' knowledge about how to combine ingredients. The complement network captures which ingredients tend to co-occur frequently, and is composed of two large communities: one savory, the other sweet. The substitute network, derived from user-generated suggestions for modifications, can be decomposed into many communities of functionally equivalent ingredients, and captures users' preference for healthier variants of a recipe. Our experiments reveal that recipe ratings can be well predicted by features derived from combinations of ingredient networks and nutrition information (with accuracy .792), while most of the prediction power comes from the ingredient networks (84\%).
The rest of the paper is organized as follows. Section~\ref{sec:relwork} reviews the related work. Section~\ref{sec:data} describes the dataset. Section~\ref{sec:complementnet} discusses the extraction of the ingredient and complement networks and their characteristics. Section~\ref{sec:modification} presents the extraction of recipe modification information, as well as the construction and characteristics of the ingredient substitute network.
Section~\ref{sec:predict} presents our experiments on recipe recommendation and Section~\ref{sec:conclusion} concludes.
\section{Related work}\label{sec:relwork}
Recipe recommendation has been the subject of much prior work. Typically the goal has been to suggest recipes to users based on their past recipe ratings~\cite{svensson2005designing}\cite{forbes2011content} or browsing/cooking history~\cite{uedauser}. The algorithms then find similar recipes based on overlapping ingredients, either treating each ingredient equally~\cite{freyne2010intelligent} or by identifying key ingredients~\cite{zhang2008back}. Instead of modeling recipes using ingredients, Wang et al.~\cite{wang2008substructure} represent the recipes as graphs which are built on ingredients and cooking directions, and they demonstrate that graph representations can be used to easily aggregate Chinese dishes by the flow of cooking steps and the sequence of added ingredients. However, their approach only models the occurrence of ingredients or cooking methods, and doesn't take into account the relationships between ingredients. In contrast, in this paper we incorporate the likelihood of ingredients to co-occur, as well as the potential of one ingredient to act as a substitute for another.
Another branch of research has focused on recommending recipes based on desired nutritional intake or promoting healthy food choices. Geleijnse et al.~\cite{geleijnse2011personalized} designed a prototype of a personalized recipe advice system, which suggests recipes to users based on their past food selections and nutrition intake. In addition to nutrition information, Kamieth et al.~\cite{kamieth2011adaptive} built a personalized recipe recommendation system based on availability of ingredients and personal nutritional needs. Shidochi et al.~\cite{shidochi2009finding} proposed an algorithm to extract replaceable ingredients from recipes in order to satisfy users' various demands, such as calorie constraints and food availability. Their method identifies substitutable ingredients by matching the cooking actions that correspond to ingredient names.
However, their assumption that substitutable ingredients are subject to the same processing methods is less direct and specific than extracting substitutions directly from user-contributed suggestions.
Ahn et al. \cite{ahn2011flavor} and Kinouchi et al \cite{kinouchi2008non} examined networks involving ingredients derived from recipes, with the former modeling ingredients by their flavor bonds, and the latter examining the relationship between ingredients and recipes. In contrast, we derive direct ingredient-ingredient networks of both compliments and substitutes. We also step beyond characterizing these networks to demonstrating that they can be used to predict which recipes will be successful.
\vspace{20pt}
\section{Dataset}\label{sec:data}
Allrecipes.com is one of the most popular recipe-sharing websites, where novice and expert cooks alike can upload and rate cooking recipes. It hosts 16 customized international sites
for users to share their recipes in their native languages, of which we study only the main, English, version. Recipes uploaded to the site contain specific instructions on how to prepare a dish: the list of ingredients, preparation steps, preparation and cook time, the number of servings produced, nutrition information, serving directions, and photos of the prepared dish. The uploaded recipes are enriched with user ratings and reviews, which comment on the quality of the recipe, and suggest changes and improvements. In addition to rating and commenting on recipes, users are able to save them as favorites or recommend them to others through a forum.
We downloaded 46,337 recipes including all information listed from allrecipes.com, including several classifications, such as a region (e.g. the midwest region of US or Europe), the course or meal the dish is appropriate for (e.g.: appetizers or breakfast), and any holidays the dish may be associated with. In order to understand users' recipe preferences, we crawled 1,976,920 reviews which include reviewers' ratings, review text, and the number of users who voted the review as useful.
\subsection{Data preprocessing}
The first step in processing the recipes is identifying the ingredients and cooking methods from the freeform text of the recipe. Usually, although not always, each ingredient is listed on a separate line. To extract the ingredients, we tried two approaches. In the first, we found the maximal match between a pre-curated list of ingredients and the text of the line. However, this missed too many ingredients, while misidentifying others. In the second approach, we used regular expression matching to remove non-ingredient terms from the line and identified the remainder as the ingredient. We removed quantifiers, such as e.g. ``1 lb'' or ``2 cups'', words referring to consistency or temperature, e.g. chopped or cold, along with a few other heuristics, such as removing content in parentheses. For example ``1 (28 ounce) can baked beans (such as Bush's Original\textregistered)" is identified as
``baked beans". By limiting the list of potential terms to remove from an ingredient entry, we erred on the side of not conflating potentially identical or highly similar ingredients, e.g. ``cheddar cheese'', used in 2450 recipes, was considered different from ``sharp cheddar cheese'', occurring in 394 recipes.
We then generated an ingredient list sorted by frequency of ingredient occurrence and selected the top 1000 common ingredient names as our finalized ingredient list. Each of the top 1000 ingredients occurred in 23 or more recipes, with plain salt making an appearance in 47.3\% of recipes. These ingredients also accounted for 94.9\% of ingredient entries in the recipe dataset. The remaining ingredients were missed either because of high specificity (e.g. yolk-free egg noodle), referencing brand names (e.g. Planters almonds), rarity (e.g. serviceberry), misspellings, or not being a food (e.g. ``nylon netting").
The remaining processing task was to identify cooking processes from the directions. We first identified all heating methods using a listing in the Wikipedia entry on cooking ~\cite{wikiCooking}. For example, baking, boiling, and steaming are all ways of heating the food. We then identified mechanical ways of processing the food such as chopping and grinding, and other chemical techniques such as marinating and brining.
\subsection{Regional preferences}
Choosing one cooking method over another appears to be a question of regional taste. 5.8\% of recipes were classified into one of five US regions: Mountain, Midwest, Northeast, South, and West Coast (including Alaska and Hawaii).
Figure~\ref{fig:region_method} shows significantly ($\chi^2$ test p-value <~0.001) varying preferences in the different US regions among 6 of the most popular cooking methods. Boiling and simmering, both involving heating food in hot liquids, are more common in the South and Midwest.
Marinating and grilling are relatively more popular in the West and Mountain regions, but in the West more grilling recipes involve seafood (18/42 = 42\%) relative to other regions combined (7/106 = 6\%). Frying is popular in the South and Northeast. Baking is a universally popular and versatile technique, which is often used for both sweet and savory dishes, and is slightly more popular in the Northeast and Midwest. Examination of individual recipes reflecting these frequencies shows that these differences in preference can be tied to differences in demographics, immigrant culture and availability of local ingredients, e.g. seafood.
\begin{figure}[t!]
\centering
\includegraphics[trim=2 20 10 10,width=1\columnwidth]{region_method.pdf}
\caption{ \label{fig:region_method}
The percentage of recipes by region that apply a specific heating method.}
\vspace{-1em}
\end{figure}
\section{Ingredient complement network}\label{sec:complementnet}
Can we learn how to combine ingredients from the data?
Here we employ the occurrences of ingredients across recipes to distill users' knowledge about combining ingredients.
We constructed an ingredient complement network based on pointwise mutual information (PMI) defined on pairs of ingredients $(a,b)$:
\[
\mathrm{PMI(a,b)} = log\frac{p(a,b)}{p(a)p(b)},
\]
where
\[
p(a,b)= \frac{\mathrm{\# \:of \:recipes\: containing} \: a \mathrm{\:and} \:b }{\mathrm{\# \:of \:recipes}},
\]
\[
p(a) = \frac{\mathrm{\# \:of \:recipes\: containing} \:a }{\mathrm{\# \:of \:recipes}},
\]
\[
p(b) = \frac{\mathrm{\# \:of \:recipes\: containing} \:b }{\mathrm{\# \:of \:recipes}}.
\]
The PMI gives the probability that two ingredients occur together against the probability that they occur separately. Complementary ingredients tend to occur together far more often than would be expected by chance.
Figure~\ref{fig:complementnet} shows a visualization of ingredient complementarity. Two distinct subcommunities of recipes are immediately apparent: one corresponding to savory dishes, the other to sweet ones. Some central ingredients, e.g. egg and salt, actually are pushed to the periphery of the network. They are so ubiquitous, that although they have many edges, they are all weak, since they don't show particular complementarity with any single group of ingredients.
\begin{figure*}[bth]
\centering
\includegraphics[trim=0 30 0 30,width=1.0\textwidth]{complements2.pdf}
\caption{Ingredient complement network. Two ingredients share an edge if they occur together more than would be expected by chance and if their pointwise mutual information exceeds a threshold. \label{fig:complementnet}}
\end{figure*}
We further probed the structure of the complementarity network by applying a network clustering algorithm~\cite{rosvall2008maps}. The algorithm confirmed the existence of two main clusters containing the vast majority of the ingredients. An interesting satellite cluster is that of mixed drink ingredients, which is evident as a constellation of small nodes located near the top of the sweet cluster in Figure~\ref{fig:complementnet}. The cluster includes the following ingredients: lime, rum, ice, orange, pineapple juice, vodka, cranberry juice, lemonade, tequila, etc.
For each recipe we recorded the minimum, average, and maximum pairwise pointwise mutual information between ingredients. The intuition is that complementary ingredients would yield higher ratings, while ingredients that don't go together would lower the average rating. We found that while the average and minimum pointwise mutual information between ingredients is uncorrelated with ratings, the maximum is very slightly positively correlated with the average rating for the recipe ($\rho = 0.09$, p-value < $10^{-10}$). This suggests that having at least two complementary ingredients very slightly boosts a recipe's prospects, but having clashing or unrelated ingredients does not seem to do harm.
\section{Recipe modifications}\label{sec:modification}
Co-occurrence of ingredients aggregated over individual recipes reveals the structure of cooking, but tells us little about how flexible the ingredient proportions are, or whether some ingredients could easily be left out or substituted. An experienced cook may know that apple sauce is a low-fat alternative to oil, or may know that nutmeg is often optional, but a novice cook may implement recipes literally, afraid that deviating from the instructions may produce poor results. While a traditional hardcopy cookbook would provide few such hints, they are plentiful in the reviews submitted by users who implemented the recipes, e.g. {\em ``This is a great recipe, but using fresh tomatoes only adds a few minutes to the prep time and makes it taste so much better"}, or another comment about the same salsa recipe {\em ``This is by far the best recipe we have ever come across. We did however change it just a little bit by adding extra onion.''}
As the examples illustrate, modifications are reported even when the user likes the recipe. In fact, we found that 60.1\% of recipe reviews contain words signaling modification, such as ``add", ``omit", ``instead'', ``extra" and 14 others. Furthermore, it is the reviews that include changes that have a statistically higher average rating (4.49 vs. 4.39, t-test p-value $< 10^{-10}$), and lower rating variance (0.82 vs. 1.05, Bartlett test p-value $< 10^{-10}$), as is evident in the distribution of ratings, shown in Fig.~\ref{fig:modornomod}. This suggests that flexibility in recipes is not necessarily a bad thing, and that reviewers who don't mention modifications are more likely to think of the recipe as perfect, or to dislike it entirely.
In the following, we describe the recipe modifications extracted from user reviews, including adjustment, deletion and addition. We then present how we constructed an ingredient substitute network based on the extracted information.
\begin{figure}[t!]
\centering
\includegraphics[trim=0 36 0 40, width=1\columnwidth]{modornot.pdf}
\caption{The likelihood that a review suggests a modification to the recipe depends on the star rating the review is assigning to the recipe. \label{fig:modornomod}}
\end{figure}
\subsection{Adjustments}
Some modifications involve increasing or decreasing the amount of an ingredient in the recipe. In this and the following analyses, we split the review on punctuation such as commas and periods. We used simple heuristics to detect when a review suggested a modification: {\em adding/using more/less} of an ingredient counted as an increase/decrease. Doubling or increasing counted as an increase, while reducing, cutting, or decreasing counted as a decrease. While it is likely that there are other expressions signaling the adjustment of ingredient quantities, using this set of terms allowed us to compare the relative rate of modification, as well as the frequency of increase vs. decrease between ingredients. The ingredients themselves were extracted by performing a maximal character match within a window following an adjustment term.
Figure~\ref{fig:updownmod} shows the ratios of the number of reviews suggesting modifications, either increases or decreases, to the number of recipes that contain the ingredient. Two patterns are immediately apparent. Ingredients that may be perceived as being unhealthy, such as fats and sugars, are, with the exception of vegetable oil and margarine, more likely to be modified, and to be decreased. On the other hand, flavor enhancers such as soy sauce, lemon juice, cinnamon, Worcestershire sauce, and toppings such as cheeses, bacon and mushrooms, are also likely to be modified; however, they tend to be added in greater, rather than lesser quantities. Combined, the patterns suggest that good-tasting but ``unhealthy" ingredients can be reduced, if desired, while spices, extracts, and toppings can be increased to taste.
\begin{figure}[t!]
\centering
\includegraphics[trim=30 40 0 30,width=1\columnwidth]{updownmod.pdf}
\caption{Suggested modifications of quantity for the 50 most common ingredients, derived from recipe reviews. The line denotes equal numbers of suggested quantity increases and decreases. \label{fig:updownmod}}
\end{figure}
\subsection{Deletions and additions}
Recipes are also frequently modified such that ingredients are omitted entirely. We looked for words indicating that the reviewer did not have an ingredient (and hence did not use it), e.g. ``had no" and ``didn't have". We further used ``omit/left out/left off/bother with'' as indication that the reviewer had omitted the ingredients, potentially for other reasons. Because reviewers often used simplified terms, e.g. ``vanilla" instead of ``vanilla extract", we compared words in proximity to the action words by constructing 4-character-grams and calculating the cosine similarity between the n-grams in the review and the list of ingredients for the recipe.
To identify additions, we simply looked for the word ``add", but omitted possible substitutions. For example, we would use ``added cucumber", but not ``added cucumber instead of green pepper", the latter of which we analyze in the following section. We then compared the addition to the list of ingredients in the recipes, and considered the addition valid only if the ingredient does not already belong in the recipe.
Table~\ref{tab:modcorr} shows the correlation between ingredient modifications. As might be expected, the more frequently an ingredient occurs in a recipe, the more times its quantity has the opportunity to be modified, as is evident in the strong correlation between the the number of recipes the ingredient occurs in and both increases and decreases recommended in reviews. However, the more common an ingredient, the more stable it appears to be. Recipe frequency is negatively correlated with deletions/recipe ($\rho = -0.22$), additions/recipe ($\rho = -0.25$), and increases/recipe ($\rho = -0.26$). For example, salt is so essential, appearing in over 21,000 recipes, that we detected only 18 reviews where it was explicitly dropped. In contrast, Worcheshire sauce, appearing in 1,542 recipes, is dropped explicitly in 148 reviews.
As might also be expected, additions are positively correlated with increases, and deletions with decreases. However, additions and deletions are very weakly negatively correlated, indicating that an ingredient that is added frequently is not necessarily omitted more frequently as well.
\begin{table}[ht!]
\caption{Correlations between ingredient modifications \label{tab:modcorr}}
\begin{center}
\begin{tabular}{lrrrr}
\hline
& addition & deletion & increase & decrease \\
\hline
\# recipes & 0.41 & 0.22 & 0.61 & 0.68 \\
addition & &-0.15 & 0.79 & 0.11 \\
deletion & & & 0.09 & 0.58 \\
increase & & & & 0.39 \\
\hline
\end{tabular}
\end{center}
\vspace{-1em}
\end{table}
\subsection{Ingredient substitute network}\label{sec:substitutenet}
Replacement relationships show whether one ingredient is preferable to another. The preference could be based on taste, availability, or price. Some ingredient substitution tables can be found online\footnote{e.g., http://allrecipes.com/HowTo/common-ingredient-substitutions/detail.aspx}, but are neither extensive nor contain information about relative frequencies of each substitution. Thus, we found an alternative source for extracting replacement relationships -- users' comments, e.g. {\em ``I replaced the butter in the frosting by sour cream, just to soothe my conscience about all the fatty calories"}.
To extract such knowledge, we first parsed the reviews as follows: we considered several phrases to signal replacement relationships: ``replace $a$ with $b$'', ``substitute $b$ for $a$'', ``$b$ instead of $a$'', etc, and matched $a$ and $b$ to our list of ingredients.
We constructed an ingredient substitute network to capture users' knowledge about ingredient replacement. This weighted, directed network consists of ingredients as nodes. We thresholded and eliminated any suggested substitutions that occurred fewer than 5 times. We then determined the weight of each edge by $p(b|a)$, the proportion of substitutions of ingredient $a$ that suggest ingredient $b$.
For example, 68\% of substitutions for white sugar were to splenda, an artificial sweetener, and hence the assigned weight for the $sugar \rightarrow splenda$ edge is 0.68.
\begin{figure}[t!]
\centering
\includegraphics[trim=10 30 0 10,width=1.2\columnwidth]{substitutes.png}
\caption{Ingredient substitute network. Nodes are sized according to the number of times they have been recommended as a substitute for another ingredient, and colored according to their indegree. \label{fig:substitutenetwork}}
\vspace{-1em}
\end{figure}
\begin{table}[ht!]
\caption{Clusters of ingredients that can be substituted for one another. A maximum of 5 additional ingredients for each cluster are listed, ordered by PageRank. \label{tab:subclusters}}
\vspace{-1em}
\begin{center}
\begin{tabular}{r|l}
\hline
main & other ingredients \\
\hline \hline
chicken & turkey, beef, sausage, chicken breast, bacon\\ \hline
olive oil & butter, apple sauce, oil, banana, margarine \\ \hline
sweet & yam, potato, pumpkin, butternut squash, \\
potato & parsnip \\ \hline
baking & baking soda, cream of tartar \\
powder & \\ \hline
almond & pecan, walnut, cashew, peanut, sunflower s. \\\hline
apple & peach, pineapple, pear, mango, pie filling \\\hline
egg & egg white, egg substitute, egg yolk \\ \hline
tilapia & cod, catfish, flounder, halibut, orange roughy \\ \hline
spinach & mushroom, broccoli, kale, carrot, zucchini \\ \hline
italian & basil, cilantro, oregano, parsley, dill \\
seasoning & \\ \hline
cabbage & coleslaw mix, sauerkraut, bok choy\\
& napa cabbage
\\ \hline
\end{tabular}
\end{center}
\vspace{-1em}
\end{table}
The resulting substitution network, shown in Figure~\ref{fig:substitutenetwork}, exhibits strong clustering. We examined this structure by applying the map generator tool by Rosvall et al.~\cite{rosvall2008maps}, which uses a random walk approach to identify clusters in weighted, directed networks. The resulting clusters, and their relationships to one another, are shown in Fig.~\ref{fig:subclusters}. The derived clusters could be used when following a relatively new recipe which may not receive many reviews, and therefore many suggestions for ingredient substitutions. If one does not have all ingredients at hand, one could examine the content of one's fridge and pantry and match it with other ingredients found in the same cluster as the ingredient called for by the recipe. Table~\ref{tab:subclusters} lists the contents of a few such sample ingredient clusters, and Fig.~\ref{fig:subcloseup} shows two example clusters extracted from the substitute network.
\begin{figure}[bth]
\centering
\includegraphics[trim=0 20 0 0,width=1\columnwidth]{substitutesinfocluster.pdf}
\caption{\label{fig:subclusters} Ingredient substitution clusters. Nodes represent clusters and edges indicate the presence of recommended substitutions that span clusters. Each cluster represents a set of related ingredients which are frequently substituted for one another.}
\end{figure}
\begin{figure}
\centering
\begin{tabular}{cc}
\includegraphics[width=0.5\columnwidth]{milksubstitutes.pdf} &
\includegraphics[width=0.5\columnwidth]{cinnamonsubstitutes.pdf}\\
(a) milk substitutes & (b) cinammon substitutes \\
\end{tabular}
\caption{Relationships between ingredients located within two of the clusters from Fig.~\ref{fig:subclusters}.} \label{fig:subcloseup}
\vspace{-1em}
\end{figure}
\pagebreak
Finally, we examine whether the substitution network encodes preferences for one ingredient over another, as evidenced by the relative ratings of similar recipes, one which contains an original ingredient, and another which implements a substitution. To test this hypothesis, we construct a ``preference network", where one ingredient is preferred to another in terms of received ratings, and is constructed by creating an edge $(a,b)$ between a pair of ingredients, where $a$ and $b$ are listed in two recipes $X$ and $Y$ respectively, if recipe ratings $R_{X}>R_{Y}$. For example, if recipe $X$ includes beef, ketchup and cheese, and recipe $Y$ contains beef and pickles, then this recipe pair contributes to two edges: one from pickles to ketchup, and the other from pickles to cheese. The aggregate edge weights are defined based on PMI. Because PMI is a symmetric quantity ($\mathrm{PMI}(a;b)=\mathrm{PMI}(b;a)$), we introduce a directed PMI measure to cope with the directionality of the preference network:
\[
\mathrm{PMI}(a\to b) = \mathrm{log}\frac{p(a\to b)}{p(a)p(b)},
\]
where
\[
p(a \to b )= \frac{\mathrm{\# \:of \:recipe\: pairs\: from}\: a \mathrm{\:to} \:b }{\mathrm{\# \:of \:recipe \:pairs}},
\]
and $p(a)$, $p(b)$ are defined as in the previous section.
We find high correlation between this preference network and the substitution network ($\rho = 0.72, p < 0.001$). This observation suggests that the substitute network encodes users' ingredient preference, which we use in the recipe prediction task described in the next section.
\section{Recipe recommendation}\label{sec:predict}
We use the above insights to uncover novel recommendation algorithms suitable for recipe recommendations. We use ingredients and the relationships encoded between them in ingredient networks as our main feature sets to predict recipe ratings, and compare them against features encoding nutrition information, as well as other baseline features such as cooking methods, and preparation and cook time. Then we apply a discriminative machine learning method, stochastic gradient boosting trees~\cite{Friedman98additivelogistic}, to predict recipe ratings.
\input{predict}
\section{Conclusion}\label{sec:conclusion}
Recipes are little more than instructions for combining and processing sets of ingredients. Individual cookbooks, even the most expansive ones, contain single recipes for each dish. The web, however, permits collaborative recipe generation and modification, with tens of thousands of recipes contributed in individual websites. We have shown how this data can be used to glean insights about regional preferences and modifiability of individual ingredients, and also how it can be used to construct two kinds of networks, one of ingredient complements, the other of ingredient substitutes. These networks encode which ingredients go well together, and which can be substituted to obtain superior results, and permit one to predict, given a pair of related recipes, which one will be more highly rated by users.
In future work, we plan to extend ingredient networks to incorporate the cooking methods as well. It would also be of interest to generate region-specific and diet-specific ratings, depending on the users' background and preferences. A whole host of user-interface features could be added for users who are interacting with recipes, whether the recipe is newly submitted, and hence unrated, or whether they are browsing a cookbook. In addition to automatically predicting a rating for the recipe, one could flag ingredients that can be omitted, ones whose quantity could be tweaked, as well as suggested additions and substitutions.
\section{Acknowledgments}
This work was supported by MURI award FA9550-08-1-0265 from the Air Force Office of Scientific Research.
The methodology used in this paper was developed with support from funding from the Army Research Office, Multi-University Research Initiative on Measuring, Understanding, and Responding to Covert Social Networks: Passive and Active Tomography. The authors gratefully acknowledge D. Lazer for support.
\bibliographystyle{acm-sigchi}
|
\section{Introduction}
Physical phenomena described by reduced or effective non-Hermitian Hamiltonians are
often encountered in a wide class of quantum or classical systems, for example in nuclear or condensed-matter physics of open systems \cite{NH1,NH2,NH3} or in optical systems
in presence of optical gain or losses \cite{NH4}. Among non-Hermitian Hamiltonians, great interest has been devoted in the past two decades to study the properties of parity-time ($\mathcal{PT}$) invariant Hamiltonians, which possess a real-valued energy spectrum below a symmetry-breaking point in spite of non-Hermiticity. Such a class of non-Hermitian Hamiltonians has been originally introduced by Carl Bender in the framework of non-Hermitian extensions of quantum mechanics and quantum field theories \cite{Bender1,Bender2,Bender3}, and found recently
an increasing interest since the proposal of physical systems described by $\mathcal{PT}$-symmetric Hamiltonians, including optical \cite{O1,O2,O3,O4,O5,O6,O7,O8,Science,refe1,refe2,refe3} and electronic \cite{electro} systems.
\par
Complex periodic potentials \cite{C1,C2,C3,C4,C4bis,C5,C6,C6tris,C6bis,C7} realize a kind of synthetic complex crystals, which show rather unusual scattering and transport properties as compared to ordinary crystals. Complex crystals
have been investigated in different areas of physics, ranging
from matter waves \cite{C4bis,C5,C6,C6tris,C6bis} to optics \cite{O3}.
In optics, a complex crystal with $\mathcal{PT}$ invariance is realized by
introduction of index and balanced gain/loss modulations in a dielectric medium \cite{O3}.
Complex crystals can be also realized in atom optics experiments exploiting
the interaction of near resonant light with an open two-level
system. Scattering of matter waves from purely absorbing optical lattices was reported in a few earlier experiments Refs.\cite{C4bis,C5,C6bis}. As noticed by Berry \cite{Berry}, although in such experiments on matter waves $\mathcal{PT}$ symmetry is not strictly realized and the complex potentials are purely absorptive, the analysis is essentially the
same, since the mean loss simply represents an overall exponential decay of the wave.
From the theoretical side, Bragg scattering, diffraction and transport properties have been extensively investigated for sinusoidal $\mathcal{PT}$-symmetric complex crystals \cite{O3,C1,C2,L1,L2,L3,L4,L5}, revealing some interesting properties such as violation of the Friedel's law of Bragg
scattering \cite{C5,C6,L1}, double refraction and nonreciprocal diffraction \cite{O3},
and unidirectional Bloch oscillations \cite{L4}. In particular, in a recent work \cite{L5} it was predicted that a sinusoidal $\mathcal{PT}$-symmetric sinusoidal crystal of finite length near the spontaneous $\mathcal{PT}$-symmetry breaking point
can act as a unidirectional invisible medium, i.e. the crystal is almost reflectionless when probed from one side, and transmission occurs as if the crystal were absent. Such an unidirectional invisibility of $\mathcal{PT}$-symmetric Bragg scatters near the symmetry-breaking point was previously predicted to occur in Ref.\cite{cg} for waveguide Bragg gratings which
combine matched periodic modulations of refractive index and loss/gain
yielding asymmetrical mode coupling. In these previous studies \cite{L5,cg}, invisibility was explained on the basis of a coupled-mode theory describing Bragg scattering and coupling of counter-propagating waves in the crystal, which is rather common in the optical context \cite{Sipe,Poladian}. Such an analysis predicts that, for a shallow grating near the $\mathcal{PT}$ symmetry breaking point, the sinusoidal crystal appears to be invisible when probed from one side {\it independently} of the crystal length. \par
In this work we re-consider the scattering properties of the sinusoidal $\mathcal{PT}$-symmetric potential and derive {\it exact} analytical expressions for reflection and transmission coefficients. The analysis shows that application of the coupled-mode theory in the standard form fails to predict the correct scattering properties in case of long crystals. In particular, as at short lengths the crystal is reflectionless and invisible when probed from one side (according to previous studies \cite{L5,cg}), at intermediate lengths the crystal remains reflectionless but not invisible. At even longer crystal lengths, both unidirectional reflectionless and invisibility properties are broken.
\section{Bragg scattering in $\mathcal{PT}$-symmetric sinusoidal potentials: general aspects and extended coupled-mode theory}
\subsection{The model}
Let us consider the stationary Schr\"{o}dinger equation for a quantum particle in a locally periodic and complex potential V(x), which in dimensionless form reads
\begin{equation}
\hat{H} \psi \equiv -\frac{d^2 \psi}{dx^2}-V(x) \psi=E \psi
\end{equation}
where $E$ is the energy of the incident particle and $V(x)$ is the complex scattering potential with period $\Lambda$, which is nonvanishing in the interval $0<x<L$. The crystal length $L$ is assumed to be an integer multiple of the lattice period $\Lambda$, i.e. $L=N \Lambda$, where $N$ is the number of unit cells in the crystal. As mentioned in the introduction, Eq.(1) describes Bragg scattering of matter waves from a complex potential in the non-interacting regime, which applies e.g. to a dilute cold atomic beam (see, for instance, \cite{C6bis}). In this case, the complex potential arises from the interaction of near resonant light with an open two-level
system, and it is generally absorptive. In this work we will mainly focus our attention to the $\mathcal{PT}$-symmetric sinusoidal potential, assuming
\begin{equation}
V(x)=V_0 \left[ \cos \left( 2 \pi x / \Lambda \right)+i \sigma \sin \left( 2 \pi x / \Lambda\right) \right]
\end{equation}
for $0<x<L$, and $V(x)=0$ for $x<0$ and $x>L$,
where $V_0$ is the lattice amplitude and $\sigma \geq 0$ measures the strength of the non-Hermitian part of the potential. The spectral properties of the $\mathcal{PT}$-symmetric sinusoidal potential (3) have been investigated in Refs.\cite{O3,C2,C7,L1,L4}. For the infinitely-extended crystal, the energy spectrum remains real-valued for $\sigma \leq 1$, and breaking of the $\mathcal{PT}$ phase is attained at $\sigma=\sigma_c=1$ \cite{O3,C2,C7,L1,L4}. Here we consider Bragg scattering of incoming waves with momentum $p$ close to the Bragg value $\pi/ \Lambda$, i.e. with energy $E=p^2$ close to $(\pi/\Lambda)^2$, and typically will assume a modulation $V_0$ of the potential much smaller than the energy $E$. \\
It should be noted that Bragg scattering of optical waves in one-dimensional Bragg grating structures, considered in Refs.\cite{L5,cg}, is basically analogous to Bragg scattering of matter waves in the framework of Eq.(1). In fact, the electric field amplitude $\mathcal{E}(x)$ of an optical wave at frequency $\omega$ that propagates along a dielectric medium with a spatially-dependent relative dielectric constant $\epsilon(x)=n_0^2[1+\Delta \epsilon (x)]$, where $n_0$ is the refractive index of the lossless medium and $\Delta \epsilon(x+\Lambda)=\Delta \epsilon(x)$ accounts for the
index and gain/loss modulation, satisfies the scalar Helmholtz equation, which can be written in the form
\begin{equation}
-\frac{d^2 \mathcal{E}}{dx^2}-E \Delta \epsilon(x) \mathcal{E}=E \mathcal{E},
\end{equation}
where we have set $E=k^2$ and $k= n_0 \omega/c_0$. Note that Eq.(3) formally reduces to the Schr\"{o}dinger equation provided that the following formal substitutions
\begin{equation}
\mathcal{E} \rightarrow \psi \; , \;\;\; k \rightarrow p
\end{equation}
are made, with a complex scattering potential $V(x)$ related to the modulation of the dielectric constant $\Delta \epsilon(x)$ by the simple relation
\begin{equation}
V(x)=E \Delta \epsilon(x).
\end{equation}
Hence, the only difference between scattering of matter waves in complex optical potentials and light waves in complex Bragg gratings is that, in the latter case, the complex scattering potential $V(x)=E\Delta \epsilon(x)$ in the equivalent Schr\"{o}dinger equation depends
on the energy $E$ of the incidence particle. However, for shallow gratings Bragg scattering occurs solely for optical fields with frequencies $\omega$ very close to the Bragg frequency $\omega_B=c_0 \pi/(n_0 \Lambda)$ (see, e.g. \cite{cg}), and thus one can safely assume $V(x) \simeq (\pi/ \Lambda)^2 \Delta \epsilon(x)$, leaving out the dependence of the scattering potential from the energy. In the following, we will mainly focus our analysis to the determination of the reflection and transmission coefficients for scattering of matter waves in the framework of Eq.(1), however similar results hold {\it mutatis mutandis} for reflection and transmission of optical waves in Bragg grating structures.
\subsection{Scattering states, spectrum, and reflection/transmission coefficients}
Since $V(x)=0$ for $x<0$ and $x>L$, the continuous spectrum of the Hamiltonian $\hat{H}$ is the semi-infinite real axis of energies $E=p^2 \geq 0$, and the corresponding eigenfunctions are the scattered states, defined by the relations
\begin{equation}
\psi(x)= \left\{
\begin{array}{c}
\alpha_1 \exp(ipx)+\beta_1 \exp(-ipx) \; \; x \leq 0 \\
\alpha_2 \exp[ip(x-L)+\beta_2 \exp[-ip(x-L)] \; \; x \geq L \\
\end{array}
\right.
\end{equation}
where $p \geq 0$ is the momentum and $(\alpha_1,\beta_1)$, $(\alpha_2,\beta_2)$ are the amplitudes of forward and backward propagating waves on the left ($x<0$) and on the right ($x>L$) sides of the crystal, respectively. Such amplitudes are related by the algebraic equation (see, for instance, \cite{O6})
\begin{equation}
\left(
\begin{array}{c}
\alpha_2 \\
\beta_2
\end{array}
\right)=\mathcal{M}(p) \left(
\begin{array}{c}
\alpha_1 \\
\beta_1
\end{array}
\right)
\end{equation}
where the $2 \times 2$ transfer matrix $\mathcal{M}(p)$ is unimodular, i.e. ${\rm det} \mathcal{M}=\mathcal{M}_{22} \mathcal{M}_{11}-\mathcal{M}_{12}\mathcal{M}_{21}=1$. For a $\mathcal{PT}$-symmetric potential, the further relation $\mathcal{M}_{22}(p)=\mathcal{M}_{11}^*(p^*)$ holds.
The transmission ($t$) and reflection
($r$) coefficients for left ($l$) and right ($r$) side incidence are related to the coefficients of the transfer matrix by the usual relations (see, for instance, \cite{O6})
\begin{equation}
t^{(l)}= \frac{1}{\mathcal{M}_{22}} , \; \; t^{(r)}=t^{(l)} \equiv t , \;\;\;\;
r^{(l)}=-\frac{\mathcal{M}_{21}}{\mathcal{M}_{22}}, \; \; r^{(r)}=\frac{\mathcal{M}_{12}}{\mathcal{M}_{22}}.
\end{equation}
Note that the transmission coefficient does not depend on the incidence side like in an ordinary crystal, whereas generally one has $|r^{(l)}| \neq |r^{(r)}|$, i.e. in reflection a complex crystal behaves differently for left and right incidence. This is a rather general result of wave scattering from a complex potential barrier which was previously discussed e.g. in \cite{ahmed,ventura}. If we indicate by $\mathcal{Z}(p)$ the fundamental matrix of Eq.(1) from $x=0$ to $z=L$ which relates the values of $\psi(x)$ and $(d \psi/dx)$ at the planes $x=0$ and $x=L$, i.e.
\begin{equation}
\left(
\begin{array}{c}
\psi(L) \\
(d \psi / dx) (L)
\end{array}
\right)=
\mathcal{Z}(p) \left(
\begin{array}{c}
\psi(0) \\
(d \psi / dx) (0)
\end{array}
\right)
\end{equation}
it can be readily shown that the transfer matrix $\mathcal{M}$ can be calculated as
\begin{equation}
\mathcal{M}(p)=\mathcal{T}^{-1}(p) \mathcal{Z}(p) \mathcal{T}(p)
\end{equation}
where we have set
\begin{equation}
\mathcal{T}(p) = \left(
\begin{array}{cc}
1 & 1 \\
ip & -ip
\end{array}
\right).
\end{equation}
From a numerical viewpoint, the fundamental matrix $ \mathcal{Z}(p)$ can be computed as follows. Let us cut the crystal into a sequence of $N_0$ thin slices, of thickness $\Delta x=L/N_0$, and let us indicate by $ \mathcal{Z}_k (p)$ the fundamental matrix associated to the propagation at the $k$-th slice ($k=1,2,...,N_0$), i.e. from $x_k=(k-1)\Delta x$ to $x_{k+1}=k \Delta x$. Then the fundamental matrix $ \mathcal{Z}(p)$
can be calculated as the ordered product
\begin{equation}
\mathcal{Z}(p)=\mathcal{Z}_N(p) \times \mathcal{Z}_{N-1}(p) \times ... \times \mathcal{Z}_2(p) \times \mathcal{Z}_1(p)
\end{equation}
If $\Delta x$ is much smaller than $\Lambda$, the potential $V(x)$ is almost constant in the interval $(x_{k},x_{k+1})$, and thus $ \mathcal{Z}_k (p)$ can be approximated as
\begin{equation}
\mathcal{Z}_k (p) \simeq \left(
\begin{array}{cc}
\cos(\lambda_k \Delta x) & (1/\lambda_k) \sin(\lambda_k \Delta x) \\
- \lambda_k \sin(\lambda_k \Delta x) & \cos(\lambda_k \Delta x)
\end{array}
\right)
\end{equation}
where we have set
\begin{equation}
\lambda_k=\sqrt{p^2+V(x_k)}.
\end{equation}
Note that, because of the periodicity of the crystal, one can limit to compute the fundamental matrix for the unit cell, $\mathcal{Z}^{(cell)}(p)$, i.e. from $x=0$ to $x= \Lambda$. The fundamental matrix of the crystal is then given by $\mathcal{Z}(p)=\mathcal{Z}^{(cell) N}(p) $, which can be computed using the relation \cite{note} $\mathcal{Z}(p)=\mathcal{Z}^{(cell)}(p) \mathcal{U}_{N_1}-\mathcal{I} \mathcal{U}_{N-2} $, where $\mathcal{I}$ is the $2 \times 2 $ identity matrix, $\mathcal{U}_N=\sin[(N+1) \theta] / \sin \theta$, and the complex angle $\theta$ is defined by $\cos \theta=(1/2) {\rm Tr} (\mathcal{Z}^{(cell)})$.
Besides scattering states, the Hamiltonian $\hat{H}$ can possess bound states belonging to the point spectrum, which are determined by the zeros of $\mathcal{M}_{22}(p)$ (i.e. the poles of the transmission coefficient $t$) in the half complex plane ${\rm Im}(p)>0$. The zeros of $\mathcal{M}_{22}(p)$ in the half complex plane ${\rm Im}(p)<0$
are resonance states. The onset of $\mathcal{PT}$ symmetry breaking is detected by the appearance of a divergence (for some real value $p=p_0 \neq 0$ of the momentum) in the transmission coefficient as the non-Hermiticity parameter $\sigma$ is increased from zero. In fact, for $\sigma=0$ the transmission and reflection coefficients are bounded from above and the zeros of $\mathcal{M}_{22}(p)$ lie in the lower half complex plane ${\rm Im}(p)<0$, i.e. they are resonances. As $\sigma$ in increased, the resonances move toward the real axis ${\rm Im}(p)=0$, until at a critical value $\sigma=\sigma_c$ a resonance crosses the real axis, say at $p=p_0 \neq 0$ (real). Correspondingly, the transmission coefficient $t(p)$ diverges at $p=p_0$. Just above $\sigma_c$, a bound state with complex energy $E=(p_0+i0^+)^2$ thus appears, which is the signature of $\mathcal{PT}$ symmetry breaking.
For a sinusoidal crystal of {\it finite} length, the transmission and reflection coefficients at $\sigma=1$ are bounded, and numerical calculations of the transfer matrix $\mathcal{M}$ (using the procedure outlined above) indicate that symmetry breaking is attained at a value $\sigma_c$ {\em larger} than one, with $\sigma_c \rightarrow 1^+$ as $L \rightarrow \infty$.
\subsection{Coupled-mode theory}
For a rather general class of $\mathcal{PT}$-symmetric complex potentials $V(x)$ describing Bragg scattering in {\it shallow} lattices, approximate expressions for the reflection and transmission coefficients can be derived by an asymptotic analysis of Eq.(1). In the optical context, such an analysis is generally referred to as the coupled-mode theory of Bragg scattering, which is known to provide accurate description of transmission and reflection coefficients for index-modulated shallow gratings (see, for instance, \cite{Sipe, Poladian}).
Such an analysis applies to Bragg scattering of particles with momentum $p$ close to $\pi / \Lambda$ (first-order Bragg scattering) provided that the particle energy $E \simeq (\pi/ \Lambda)^2$ is much larger than the characteristic modulation depth $V_0$ of the complex crystal. For the sinusoidal crystal defined by Eq.(2), this means that the parameter $\alpha=\Lambda^2 V_0/ \pi^2$ should be much smaller than one. In Ref.\cite{L5,cg}, it was shown that application of coupled-mode theory to the sinusoidal $\mathcal{PT}$-symmetric crystal
at $\sigma=1$ gives the following expressions for the transmission and reflection coefficients
\begin{equation}
t(p)= \exp(ipL) , \;\; r^{(l)}(p) = 0 , \;\; r^{(r)}(p)=\frac{i V_0 \Lambda}{2 \pi} \frac{\sin(\delta L)}{\delta} \exp[i(p+\pi/ \Lambda)L ] \;\;\;\;
\end{equation}
where we have set $\delta=p-\pi/\Lambda$.
Such equations clearly indicate that, for left-side incidence, the crystal appears to be fully invisible, i.e. there are not reflected waves and the transmitted ones propagate as if the crystal were absent \cite{L5}. Conversely, for right-side incidence as the transmitted wave propagates again as if the crystal were absent, a reflected wave is generated, with a reflectance $R^{(R)}=|r^{(l)}|^2$ that grows quadratically with the crystal thickness $L$ at Bragg resonance $\delta=0$, i.e. for $p=\pi/\Lambda$ . Such a physically relevant behavior was referred to as {\em unidirectional invisibility} in Ref.\cite{L5}. As it is shown in the Appendix A and briefly mentioned in Ref.\cite{L5}, in the framework of the coupled-mode theory unidirectional invisibility is predicted to occur for a rather general class of complex potentials with zero mean, $V(x)= \sum_{n \neq 0} \Phi_n \exp(2 \pi i n x/ \Lambda)$, provided that the condition $\Phi_{-1}=0$ (or similarly $\Phi_{1}=0$) is satisfied. Note that the $\mathcal{PT}$-symmetric sinusoidal potential (2) at $\sigma=1$ belongs to such a general class of complex potentials.\\
In the next section, we will derive {\em exact} expressions for transmission and reflection coefficients for the $\mathcal{PT}$-symmetric sinusoidal crystal, and will show that the invisibility property of the crystal, as predicted by Eq.(15), occurs solely for short crystals, and breaks down for long crystals. Here we discuss the reasons of failure of Eq.(15) to predict the correct expressions of transmission and reflection coefficients in long crystals, and propose an {\em extended} version of coupled-mode theory to properly describe Bragg scattering in complex crystals.
The derivation of coupled-mode equations is routinely done by means of averaging or multiple-scale asymptotic techniques, which is detailed in the Appendix A for the general case of a complex potential $V(x)=\sum_{n \neq 0} V_n \exp(2 i \pi n x / \Lambda)$ with zero mean. Here we give explicit analytical results for the sinusoidal $\mathcal{PT}$-symmetric crystal at $\sigma=1$, however as shown in the Appendix A similar results are obtained for a more general complex crystal provided that the condition $\Phi_{-1}=0$ is satisfied. For a small value of $\alpha$, a solution to Eq.(1) can be searched as a power series expansion
\begin{equation}
\psi(x)=\psi^{(0)}(x)+ \alpha \psi^{(1)}(x)+\alpha \psi^{(2)}(x)+...
\end{equation}
and multiple spatial scales $X_0=x$, $X_1= \alpha x$, ... are introduced to satisfy solvability conditions at the various orders in the asymptotic analysis.
If the analysis is pushed up to the order $\sim \alpha$, the solution to Eq.(1) inside the crystal for a small value of $\delta=p-\pi/\Lambda$ (of order $ \sim \alpha$) can be written as
\begin{equation}
\psi(x)=\psi^{(0)}(x)+ \alpha \psi^{(1)}(x)+o(\alpha^2)
\end{equation}
where we have set
\begin{eqnarray}
\psi^{(0)}(x) & = & u(x) \exp(i \pi x/\Lambda)+v(x) \exp(-i \pi x / \Lambda) \\
\alpha \psi^{(1)}(x) & = & \frac{V_0 \Lambda^2 u(x)}{8 \pi^2} \exp(3 i \pi x / \Lambda)
\end{eqnarray}
and where the amplitudes $u$ and $v$ satisfy the coupled-mode equations
\begin{eqnarray}
i \frac{du}{dx} & = & -\delta u-\frac{V_0 \Lambda}{2 \pi}v \\
i \frac{dv}{dx} & = & \delta v
\end{eqnarray}
From Eqs.(20) and (21), it follows that the amplitudes $u$ and $v$ at the planes $x=0$ and $z=L$ are related by the relation $(u(L),v(L))^T=\mathcal{K}(p) (u(0),v(0))^T$, where the matrix $\mathcal{K}(p)$ reads explicitly
\begin{equation}
\mathcal{K}(p)= \left(
\begin{array}{cc}
\exp(i \delta L) & i \frac{V_0 \Lambda}{2 \pi} \frac{\sin (\delta L)}{\delta} \\
0 & \exp(-i \delta L)
\end{array}
\right).
\end{equation}
In standard coupled-mode theory \cite{Sipe,Poladian}, only the leading order term $\psi^{(0)}(x)$ in the expansion (17) is considered for the computation of the transfer matrix $\mathcal{M}$, and one simply has
\begin{equation}
\mathcal{M}(p)=\mathcal{S}(p) \mathcal{K}(p)
\end{equation}
where we have set
\begin{equation}
\mathcal{S}(p)=\left(
\begin{array}{cc}
\exp(i \pi L / \Lambda) & 0 \\
0 & \exp(-i \pi L / \Lambda)
\end{array}
\right).
\end{equation}
Using Eq.(23) for the expression of the transfer matrix, one then obtains Eq.(15) for the reflection and transmission coefficients. However, for a complex crystal such an analysis can fail even if the correction term $\alpha \psi^{(1)}$, given by Eq.(19), remains smaller than $\psi^{(0)}$. To understand why this may happen, let us consider as an example the case of left-side incidence. This case requires the absence of incident waves from the right side, which formally implies $(d \psi/dx)=ip \psi$ at $x=L$, where $p$ is the momentum of the incident wave from the left side. If the approximation $\psi(x) \simeq \psi^{(0)}(x)$ is made, the appropriate boundary conditions would be $u(0)=1$ and $v(L)=0$. The key point is that, if the expression (17) of $\psi(x)$ up to the order $\sim \alpha$ is now considered, the boundary condition $v(L)=0$ does not exactly correspond to the absence of incident waves from the right side, just because of the (small) additional contribution to $\psi(x)$ given by $\alpha \psi^{(1)}(x)$ [Eq.(19)]. Hence a more accurate procedure should use the boundary conditions $u(0)=1$ and $v(L)=\epsilon$, where $\epsilon$ is a small parameter (of order $\alpha$) to be determined such that $(d \psi/dx)=ip \psi$ at $x=L$. It is obvious that, if the solution to the coupled-mode equations (20-21) with the boundary conditions $u(0)= 0$ and $v(0)=\epsilon$ (corresponding to probing the crystal from the {\em right} side with a small amplitude $\epsilon$) would remain small uniformly in the interval $(0,L)$, the additional contribution to the solution arising from taking $\epsilon \neq 0$ would just introduce a small correction to the solution corresponding to the boundary conditions $u(0)= 1$ and $v(0)=0$. Hence a small correction to the transmission and reflection coefficients Eq.(15) would be obtained. This case always occurs for short enough crystals. However, if the crystal length $L$ is long enough such that the reflection $r^{(r)}$ for right-side incidence becomes large (of the order or larger than $ \sim 1 / \alpha$), then the correction arising from the solution to the coupled-mode equations (20-21) with the boundary conditions $u(0)= 0$ and $v(0)=\epsilon$ can not be neglected anymore, and thus should be accounted for when calculating the reflection and transmission coefficients with the appropriate boundary conditions. The crystal length $L$ at which failure of Eq.(15) is expected to occur can be estimated by imposing $|r^{(r)}| \simeq 1 / \alpha$. Taking for $|r^{(r)}|$ its peak value at $\delta=0$, i.e. $|r^{(r)}| \sim V_0 \Lambda L /(2 \pi)$, one then expects failure of Eq.(15) for $L \geq \sim L_c$, where
\begin{equation}
L_c = \frac{2 \pi^3}{V_0^2 \Lambda^3}.
\end{equation}
Hence, for $L$ of the order of larger than $L_c$, Eq.(15) can not be used to calculate the transmission and reflection coefficients of the crystal. A more appropriate procedure to compute the transfer matrix, and thus the transmission and reflection coefficients, is to use Eq.(10), in which the fundamental matrix $\mathcal{Z}$ is calculated using for $\psi(x)$ the expression given by Eqs.(17-19), i.e. {\em including} the first-order correction term $\alpha \psi^{(1)}(x)$. Such an extension of the ordinary coupled-mode theory will be referred to as the {\it extended} coupled-mode theory.
\section{Bragg scattering in sinusoidal $\mathcal{PT}$-symmetric crystals: Exact analysis}
Exact expressions for the reflection and transmission coefficients can be derived for the sinusoidal potential (2) at $\sigma=1$ in terms of modified Bessel functions of firts kind. In fact, after the change of variable $y= ( \Lambda \sqrt{V_0}/ \pi) \exp(i \pi x / \Lambda)$, Eq.(1) reduces to the Bessel equation \cite{L4}
\begin{equation}
y^2 \frac{d^2 \psi}{dy^2}+y \frac{d \psi}{dy}-(y^2+q^2) \psi=0
\end{equation}
where we have set
\begin{equation}
q= \frac{p \Lambda}{\pi}.
\end{equation}
Note that Bragg scattering for particles with momentum $p$ close to $ \pi / \Lambda$ corresponds to $q \sim 1$. For $ q \neq 1$, two linearly independent solutions to Eq.(26) are $I_q(y)$ and $I_{-q}(y)$, where $I_q(y)$ is the modified Bessel function $I$ of first kind \cite{Abramowitz}. Hence in the interval $0<x<L$ two linearly independent solutions to Eq.(1) are given by
\begin{equation}
\Phi_1(x)=I_q(\Delta \exp(i \pi x/\Lambda)) \; , \;\; \Phi_2(x)=I_{-q}(\Delta \exp(i \pi x/\Lambda))
\end{equation}
where we have set
\begin{equation}
\Delta=\frac{\Lambda \sqrt{V_0}}{ \pi}= \sqrt{\alpha}.
\end{equation}
Using Eq.(28), one can construct the fundamental matrix ${\mathcal Z}(x)$ of Eq.(1) from $x=0$ to $x=L$ as
\begin{equation}
\mathcal{Z}(p) = \left(
\begin{array}{cc}
\Phi_1(L) & \Phi_2(L) \\
\Phi_1^{'}(L) & \Phi_2^{'}(L)
\end{array}
\right) \times
\left(
\begin{array}{cc}
\Phi_1(0) & \Phi_2(0) \\
\Phi_1^{'}(0) & \Phi_2^{'}(0)
\end{array}
\right) ^{-1}
\end{equation}
where the apex denotes the derivative with respect to $x$. The transfer matrix $\mathcal{M}(p)$ is finally obtained after substitution of Eq.(30) into Eq.(10). After some lengthy calculations, which are briefly detailed in the Appendix B, the following expressions for the transfer matrix coefficients are obtained:
\begin{eqnarray}
\mathcal{M}_{11}(p) = \cos(pL)+ i \frac{\Lambda \sin (pL) }{2p \sin( \pi q)} \left( p^2 Q_1 Q_2- V_0 D_1 D_2 \right) \\
\mathcal{M}_{12}(p) = - i \frac{\Lambda \sin (pL) }{2p \sin( \pi q)} \left[ V_0 D_1 D_2 +p^2 Q_1 Q_2 +p \sqrt{ V_0} \left( D_1Q_2+D_2Q_1 \right) \right] \\
\mathcal{M}_{21}(p) = i \frac{\Lambda \sin (pL) }{2p \sin( \pi q)} \left[ V_0 D_1 D_2 +p^2 Q_1 Q_2 -p \sqrt{V_0} \left( D_1Q_2+D_2Q_1 \right) \right] \\
\mathcal{M}_{22}(p) = \cos(pL)- i \frac{\Lambda \sin (pL) }{2p \sin( \pi q)} \left( p^2 Q_1 Q_2- V_0 D_1 D_2 \right)
\end{eqnarray}
where we have set
\begin{equation}
Q_1=I_q(\Delta) \; , \; Q_2=I_{-q}(\Delta) \; , \; D_1=I^{'}_q(\Delta) \; , \; D_2=I_{-q}^{'}(\Delta)
\end{equation}
and where $q$ and $\Delta$ are defined by Eqs.(27) and (29), respectively. The reflection and transmission coefficients $r^{(l,r)}(p)$ and $t(p)$ are then obtained after substitution of Eqs.(31-34) into Eq.(8). In particular, for the transmission coefficient $t(p)$ one obtains explicitly
\begin{equation}
t(p)=\frac{1}{\cos(pL)-iF(p) \sin(pL)}
\end{equation}
where we have set
\begin{equation}
F(p)=\frac{\Lambda}{2p \sin( \pi q)} \left(p^2Q_1Q_2-V_0D_1D_2 \right).
\end{equation}
Note that Eq.(36) would reduce to the first of the Eq.(15), corresponding to unidirectional crystal invisibility, if the function $F(p)$ were replaced by $1$.\\
The previous equations (31-37) have been derived for Bragg scattering of matter waves in the framework of the Schr\"{o}dinger equation (1), however similar relations hold for Bragg scattering of light waves in a complex Bragg grating structure, governed by the similar equation (3). Specifically, in view of Eqs.(4) and (5), for a grating structure with a sinusoidal $\mathcal{PT}$-symmetric modulation of the complex relative dielectric constant $\Delta \epsilon(x)= \Phi \exp(2 i \pi x / \Lambda)$ and for incident light waves with frequency $\omega$, the analytical expressions given above still hold, provided that the following formal substitutions $ p \rightarrow \omega n_0 /c_0$ and $V_0 \rightarrow (n_0 \omega/c_0)^2 \Phi \simeq (\pi/\Lambda)^2 \Phi$ are made, where $n_0$ is the refractive index of the lossless dielectric medium and $c_0$ the speed of light in vacuum.
\section{Unidirectional crystal invisibility}
Let us know discuss the unidirectional invisibility of the sinusoidal $\mathcal{PT}$-symmetric crystal on the basis of the exact scattering results presented in the previous section. To study the exact behavior of $t(p)$ as given by Eq.(36) and breakdown of crystal transparency as the number of cells $N$ is increased, let us measure the length $x$ in units of $\Lambda / \pi$, i.e. let us set without loss of generality $\Lambda=\pi$. With such a scaling, one has $\alpha=V_0$, $q=p$, $L= N \pi$ and $\Delta= \sqrt{V_0}$. Using the identity $I^{'}_p (\Delta)=I_{p-1}(\Delta)-(p/\Delta) I_{p}(\Delta)=I_{p+1}(\Delta)+(p/\Delta) I_{p}(\Delta)$ for the derivative of modified Bessel functions \cite{Abramowitz}, one can write
\begin{equation}
t(p)=\frac{1}{\cos(N \pi p)-i
F(p) \sin(N \pi p)}
\end{equation}
with
\begin{equation}
F(p)=\frac{\pi}{2p \sin( \pi p)} \left[ \sqrt{V_0} p (I_{-p} I_{p-1}+ I_pI_{-p+1})-V_0 I_{p-1}I_{-p+1} \right]
\end{equation}
\begin{figure}[htb]
\centerline{\includegraphics[width=10cm]{Fig1}} \caption{
Behavior of (a) transmittance, (b) reflectance for left-side incidence, (c) normalized phase time (in transmission), and (d) reflectance for right-side incidence versus momentum $p$ of the incident particle in a sinusoidal $\mathcal{PT}$-symmetric crystal for parameter values $\Lambda= \pi$, $V_0=0.02$ and for a number of cells $N=50$. The curves are obtained by using the exact expressions Eqs.(31-34) for the transfer matrix coefficients.}
\end{figure}
\begin{figure}[htb]
\centerline{\includegraphics[width=10cm]{Fig2}} \caption{
Same as Fig.1, but for $N=2000$. In the figures, solid lines refer to the exact scattering results obtained from Eqs.(31-34), whereas the dashed curves are the predictions based on the extended coupled-mode theory.}
\end{figure}
and where the Bessel functions are calculated at $\sqrt{V_0}$ . Note that $t(p)$ depends on two parameters solely: the number of crystal cells $N$ and the potential amplitude $V_0$. Note also that Eq.(38) is valid regardless of the smallness of $V_0$ and far from the Bragg resonance condition $p=1$ as well. Similar expressions can be derived for the reflection coefficients $r^{(l,r)}(p)$ in terms of modified Bessel functions. In the computation of the reflection and transmission coefficients, the modified Bessel functions have been calculated using a fast and highly accurate routine, discussed in Ref. \cite{Amos}. The accuracy of our procedure has been tested by checking the agreement of the results obtained from the exact analytical prediction [Eqs. (38) and (39)] and from the full numerical procedure outlined in Sec.2.2 [Eqs.(10-14)]. \par
Figures 1 and 2 show the behaviors of spectral transmittance $T(p)=|t(p)|^2$ and reflectances $R^{(l,r)}(p)=|r^{(l,r)}(p)|^2$ for left and right side incidence, as calculated by the exact analysis (solid curves), for increasing values of the number of cells $N$ and for $V_0=0.02$. In the figures, the behavior of the normalized phase time of transmitted waves, defined by $\tau_t(p)=(1/L)(d \phi_t / dp)$, is also depicted, where $\phi_t(p)$ is the phase of $t(p)$. Physically, $\tau_t(p)$ represents the traversal time of a narrow wave packet, with central momentum $p$, across the crystal, normalized to the transit time in vacuum (i.e. in the absence of the crystal). For a relatively small numbers of cells, as in Fig.1, unidirectional invisibility is observed, and reflection for left-side crystal incidence is extremely small, according to the coupled-mode theory of Sec.2.3 and the results of Refs.\cite{L5,cg}. However, as the number of cells is increased to become comparable or larger than $N_c=L_c/ \Lambda$, given by [see Eq.(25)]
\begin{equation}
N_c \sim \frac{L_c}{\Lambda}=\frac{2}{\pi \alpha^2},
\end{equation}
the invisibility regime breaks down near $p=1$, with the appearance of oscillations of the transmittance and phase time, as one can clearly see in Figs.2(a) and (c). In the figures, the predictions of spectral transmittance and reflectance computed by the {\em extended} coupled-mode theory, discussed at the end of Sec.2.3, are also depicted by the dotted curves. Note that, as the crystal is not anymore transparent, the reflectance for left-side incidence remains extremely small [see Fig. 2(b)]. In such a regime the crystal is not invisible, however it is still unidirectional {\em reflectionless}. As the number of cells is further increased, the transmittance grows in a narrow interval near the Bragg condition $p \simeq 1$, as well the the reflectance for right-side incidence. As the transmittance becomes large enough, such that $\mathcal{M}_{22}$ becomes smaller and of the same order of magnitude than $\mathcal{M}_{21}$, a very narrow resonance peak appears in the reflectance spectrum for left-side incidence. This is shown in Fig.3, in which solid and dotted curves refer to the exact results and to the approximate ones based on the extended coupled-mode theory, respectively. In such a regime, both unidirectional invisibility and reflectionless properties of the complex crystal are thus broken.To estimate the number of cells $N_c^{'}>N_c$ at which such a second transition occurs, we can apply the extended coupled-mode theory, discussed in Sec.2.3, to calculate an approximate expression of the reflection coefficient $r^{(l)}$ for left-side incidence. Using Eqs.(17-21) for an approximate expression of $\psi(x)$ and applying the appropriate boundary conditions, corresponding to left-side incidence, at exact Bragg resonance ($p=\pi/\Lambda$) an approximate expression for $|r^{(l)}|$ can be derived, which reads explicitly $r^{(l)} \sim (\pi / 64) \alpha^3 (L/ \Lambda)$. The critical number of cells $N_{c}^{'}$ at which the crystal is not anymore reflectionless for left-side incidence can be estimated by letting $|r^{(l)}| \sim 1$, which yields
\begin{equation}
N_c^{'} \sim \frac{64}{\pi \alpha^3} .
\end{equation}
\begin{figure}[htb]
\centerline{\includegraphics[width=10cm]{Fig3}} \caption{
Transmittance (left panels) and reflectance for left-side incidence (right panels) in a sinusoidal $\mathcal{PT}$-symmetric crystal for $V_0=0.02$, $\Lambda= \pi$ and for increasing number of crystal cells: (a) $N=10000$, (b), $50000$, and (c) $N=1600000$.}
\end{figure}
\section{Conclusion}
In this work Bragg scattering in sinusoidal $\mathcal{PT}$-symmetric complex crystals of finite thickness has been theoretically investigated, and exact analytical expressions for reflection and transmission coefficients have been derived in terms of
modified Bessel functions. The analytical results indicate that unidirectional invisibility, recently predicted for such crystals by coupled-mode theory \cite{L5,cg},
breaks down for crystals containing a large number of unit cells. In particular, for a given modulation depth in a shallow sinusoidal potential, three regimes have been found as the crystal length is increased. At short lengths the crystal is reflectionless and invisible when probed from one side (unidirectional invisibility), according to standard coupled-mode theory. As the numbers of cells is increased, absence of reflection for one side incidence is still observed, however the crystal is no more invisible because large oscillations in the transmittance and in the transmission phase time appear near the Bragg resonance. For still thicker crystals, both unidirectional reflectionless and invisibility properties are broken.
An extension of coupled-mode theory has been proposed to properly modelling the scattering properties of complex crystals.
|
\section{Introduction}
Knot spaces have recently been the subject of much interest. Let $\Emb(\hat{S}^1,\R^d)$ be the space of embeddings from $S^1$ to $\R^d$ with fixed initial point and initial tangent vector, which is homotopy equivalent to the space of long knots. Using Goodwillie-Weiss embedding calculus, Sinha \cite{Sinh02} defines spectral sequences converging to the homology and cohomology of $\Emb(\hat{S}^1,\R^d)$ for $d>3$. Lambrechts, Turchin and Volic \cite{LTV06} have shown that the rational cohomology spectral sequence collapses at the $E_2$ page. There is another spectral sequence, due to Vassiliev \cite{Vass92}, which converges to the homology of $\Emb(S^1,\R^d)$. The $E_1$ term of Vassiliev's spectral sequence agrees with the $E_2$ term of the embedding calculus spectral sequence by work of Turchin \cite{Tour07}. These approaches allow one to combinatorially understand the ranks of the homology groups of $\Emb(S^1,\R^d)$, but do not immediately give geometric understanding or representing cycles and cocycles in knot spaces. We present representing cycles and cocycles defined through techniques which apply to all classes in the spectral sequence.
In \cite{CCL02} Cattaneo, Cotta-Ramusino and Longoni produce explicit, nontrivial, $k(d-3)-$dimensional cycles and cocycles. We give a brief summary of these results in Section 3. They define a chain map from a graph complex to the de Rham complex of $\Emb(S^1,\R^d)$, and produce cocycles as images of graph cocycles consisting of trivalent graphs. To produce cycles, they use families of resolutions of singular knots with $k$ transverse double points. These cycles all live along the $(-2q,q(d-1))-$diagonal in the first page of the homology spectral sequence, which also serves as a vanishing line. To establish nontriviality, they show the pairing between certain cycles and cocycles is nonzero. For $d$ odd, Sakai produces a $(3d-8)-$dimensional cocycle in the space of long knots coming from a non-trivalent graph cocycle. To establish the nontriviality of this cocycle, he evaluates it on a cycle produced using the Browder bracket coming from the action of the little two-cubes operad on the space of framed knots.
The main result of this paper is the explicit production of a nontrivial cycle which lives off of the vanishing line of the homology spectral sequence for $d$ even, using techniques which should generalize. We define this cycle by generalizing the methods of Cattaneo, Cotta-Ramusino and Longoni to families of resolutions of singular knots with triple points. In particular, we first define a topological manifold $M_\beta$ and an embedding of $M_\beta$ into $\Emb(\hat{S}^1,\R^d)$, extending and correcting the results in a preprint of Longoni \cite{Long04}. Longoni also defines a cocycle which is the image of a non-trivalent graph when $d$ is even. We show that the pairing between Longoni's cocycle and our cycle is nonzero and thus both are nontrivial.
Our cycle generalizes, and our techniques are closely related to the spectral sequence combinatorics, giving possible recipes for representatives of all cycles in the embedding calculus spectral sequence. This is in contrast to Sakai's approach, which would require new input for any Browder-primitive classes off of the $(-2q,q(d-1))-$diagonal. These results will appear in future work, but we discuss them briefly at the end of this paper.
\section{Definition of the cycle}\label{sec:cycledef}
The idea at the heart of our method to produce homology classes in knot spaces goes back to Vassiliev's seminal work \cite{Vass92}. In finite type knot theory, one defines the derivative of a knot invariant by taking an immersion with transverse double-points and evaluating the knot invariant on the resolutions of that immersion. We require a generalization of such immersions.
\begin{definition}
An immersion $\gamma:S^{1} \hookrightarrow \R^{d}$ has a transverse intersection $r$-singularity at $\bar{t} = (t_1,t_2,\ldots,t_r)\in \I^{\times r}$ with $0< t_1<t_2<\cdots <t_r<1$, if all of the $\gamma(t_{i})$ coincide and the derivatives $\gamma'(t_{i})$ are generic in the sense that any $d$ or fewer of them are linearly independent.
\end{definition}
To connect with the language naturally produced by the embedding calculus spectral sequence, we use bracket expressions to encode singularity data. Sinha calculates in \cite{Sinh02} that the
subgroup of $\mathcal{P}ois^d(p)$, the $p-$th entry of the Poisson operad (see \cite{Sinh06.3}), generated by expressions with $q$ brackets such that each $x_i$ appears inside a bracket pair and the multiplication $``\cdot"$ does not appear inside a bracket pair, is also a subgroup of $E^1_{-p,q(d-1)}$ in the reduced homology spectral sequence. This is the full $E^1_{-p,q(d-1)}$ in the spectral sequence converging to the homology of the space of embeddings modulo immersions. On this subgroup, the differential $d_1: E^1_{-p,q(d-1)}\rightarrow E^1_{-p-1,q(d-1)}$ is $d^1=\sum_{i=0}^p (-1)^i(\delta^i)_*$, where $(\delta^0)_*$ is defined by adding $x_1$ in front of the expression and replacing each $x_j$ by $x_{j+1}$, $(\delta^{p+1})_*$ is defined by adding $x_{p+1}$ to the end, and for $1\leq i\leq p$, the map $(\delta^i)_*$ is defined by replacing $x_i$ by $x_i\cdot x_{i+1}$ and $x_j$ by $x_{j+1}$ for $j>i$. In \cite{Tour07}, Tourtchine does further calculations in this spectral sequence.
\begin{example} The bracket expression $\beta= \beta_1 +\beta_2$ where $\beta_1= \left[ [x_1,x_4],x_3\right] \cdot [x_2,x_5 ]$ and $ \beta_2= [x_1,x_4]\cdot \left[ [x_2,x_5],x_3\right]$ is a cycle in $E^1_{-5,3(d-1)}$.
\end{example}
\begin{definition} A pair $(\gamma, \bar{t})$ of an immersion and a sequence $\bar{t} = 0<t_1< t_2< \cdots < t_p<1$ respects a bracket expression $\beta \in \mathcal{P}ois^d(p)$ if $\gamma$ has a transverse $r$-singularity at the sequence $0<t_{i_1}< \ldots< t_{i_r}<1$ whenever $x_{i_1},\ldots,x_{i_r}$ appear inside of a bracket in $\beta$.
\end{definition}
For example, the knots $K_1$ and $K_2$ in Figure ~\ref{fig:K1K2} respect $\beta_1$ and $\beta_2$, respectively. A knot can respect a bracket expression but have higher singularities; for example $K_1$ also respects $[x_1,x_3]\cdot [x_2,x_4]$.
\begin{definition}We will denote the subspace of all pairs $(\gamma, \bar{t})\in \Imm(\hat{S}^1,\R^d)\times \I^{\times r}$ respecting a bracket expression by $\Imm_{\geq \beta}(\hat{S}^1,\R^d)$, with the convention $\Imm_\phi(\hat{S^1},\R^d) = \Imm(\hat{S^1},\R^d)$. The subspace of $\Imm_{\geq \beta}(\hat{S}^1,\R^d)$ consisting of immersions which do not have higher singularities will be denoted by $\Imm_{=\beta} (\hat{S}^1,\R^d)$. \end{definition}
\begin{figure}[h]
\centering
$$ \includegraphics[width=3in]{K1K2.png} $$
\caption{The singular knots $K_1$ and $K_2$.}
\label{fig:K1K2}
\end{figure}
In the spectral sequence, bracket expressions of the form $\prod_{m=1}^k [x_{i_m},x_{j_m}] $ are $E^1$-cycles. Submanifolds representing these cycles are well known and described in Section 2 of \cite{CCL02}. Briefly, we start with a singular knot $K\subset \R^d$ with $k$ double points which respects $\prod_{m=1}^k [x_{i_m},x_{j_m}]$, and resolve each double point by moving one strand passing through the double point off of the other. For each vector in $S^{d-3}$ we have a possible direction in which to move the strand, and therefore a possible way to resolve the double point. The subset of $\Emb(\hat{S}^1,\R^d)$ consisting of all such resolutions of $K$ is a submanifold parameterized by $\prod_m S^{d-3}$, and its fundamental class corresponds to the cycle $\prod_{m=1}^k [x_{i_m},x_{j_m}] $ of the spectral sequence.
For higher singularities, we start with ideas of Longoni \cite{Long04} and produce resolutions of transverse intersection singularities by moving one strand at a time off the intersection point. Assume the rank of the singularity $r$ is less than $d$, so the (tangent vectors of the) strands in question span a proper subspace. There are two cases - resolving a double point and resolving a higher singularity. If $r\geq 3$, we are moving a strand off the intersection point. The complementary subspace to the (tangent vector of the) strand has a unit sphere $S^{d-2}$ which parametrizes the directions to move one strand off the intersection point. If $r = 2$, we consider a unit sphere $S^{d-3}$ in the complimentary subspace which parametrizes the directions to move one strand off another.
Resolutions of triple point singularities (and higher singularities) can produce further singularities (see Figure \ref{fig:K3to6}). By restricting away from neighborhoods of those ``additional singularity'' resolutions, we produce submanifolds with boundary which we show can be pieced together to build representatives of $E^1$-cycles in the spectral sequence. We formalize as follows.
\begin{definition}
If $\beta$ is a bracket expression, let $\beta(\hat{i})$ denote the bracket expression obtained from $\beta$ by removing $x_i$ and the minimal set of other symbols as required to have a bracket expression, and replacing $x_k$ by $x_{k-1}$ for all $k>i$.
\end{definition}
For example, with $\beta_1= \left[ [x_1,x_4],x_3\right] \cdot [x_2,x_5 ]$, we have $\beta_1(\hat{4}) = [x_1,x_3][x_2,x_4]$. For each strand through a transverse intersection $r$-singularity, we can define a resolution map which moves that strand off of the singularity. To accommodate the two cases, we let \[d(r) = \left\{ \begin{array}{cc} d-2 & \mathrm{if}\; r> 2\\ d-3 &\mathrm{if}\; r= 2\end{array}\right..\]
By the rank of $x_i$ in a bracket expression $\beta$, we will mean the number of variables in $\beta$ (counting $x_i$) which appear inside of common brackets with $x_i$. In $\beta_1$, $x_3$ has rank three and $x_5$ has rank two.
\begin{definition}
If $\beta$ is a bracket expression in which $x_i$ has rank $r$ (with $r > 0$)
define the resolution map \[\rho_{i} : \Imm_{\geq \beta}(\hat{S}^1,\R^d) \times S^{d(r)} \times \I \times \I
\to \Imm_{\geq \beta(\hat{i})}(\hat{S}^1,\R^d)\] by \[\rho_i(\gamma,\bar{t},v,a,\varepsilon)(t) = \left\{ \begin{array}{cc} \gamma(t) + a\cdot v \exp\left( \frac{1}{(t-t_i)^2-\varepsilon^2}\right) & \mathrm{if}\; t\in(t_i-\varepsilon,t_i+\varepsilon)\\ \gamma(t) & \mathrm{otherwise} \end{array} \right. .\] \end{definition}
We call the triple $(v, a, \varepsilon) \in S^{d(r)} \times \I \times \I$ the resolution data. We often fix $a$ and $\varepsilon$ so that the resolutions do not have unexpected singularities and by abuse denote the restriction by $\rho_{i}$ as well. The resolution map produces immersions in which the strand (between times $t_i - \ep$ and $t_i + \ep$) is moved in the direction of $v$, as shown in Figure ~\ref{fig:resolution}.
\begin{figure}[h]
\centering
$$ \includegraphics[width=2.2in]{resolution.png} $$
\caption{The resolution of a double point.}
\label{fig:resolution}
\end{figure}
\begin{definition}\label{def:resmap}
Let $S = \{ x_{i_1},x_{i_2},\ldots,x_{i_k}\}$ be an ordered subset of the variables in $\beta$. Define $\rho_{\beta, S}$ to be the composite
$$\rho_{i_{k}} \circ (\rho_{i_{k-1}} \times id) \circ \cdots \circ (\rho_{i_{1}} \times id) :
\Imm_{\geq \beta}(\hat{S}^1,\R^d) \times \prod_{m} \left(S^{d(r_{m}) }\times \I \times \I\right) \to \Imm_{\geq \varnothing}(\hat{S}^1,\R^d),$$
where $r_{m}$ is the rank of $x_{i_{m}}$ in $\beta(\hat{i_{1}}, \ldots, \hat{i}_{m-1})$.
\end{definition}
The set $S$ encodes which strands get moved in the resolution defined by $\rho_{\beta,S}$.
We now specialize. Let $\beta_{1} = \left[ [x_1,x_4],x_3\right] \cdot [x_2,x_5 ]$, $ \beta_2= [x_1,x_4]\cdot \left[ [x_2,x_5],x_3\right] $ and choose the ordered subset of variables for each to be $S=\{x_3,x_4,x_5\}$. We choose embeddings $K_1$ and $K_2$ of $S^1$ in $\R^3\hookrightarrow \R^d$ as shown in Figure ~\ref{fig:K1K2}, as well as a sequence $0<t_1<t_2<\cdots<t_5<1$ so that $(K_1,\bar{t})$ respects $\beta_1$ and $(K_2,\bar{t})$ respects $\beta_2$.
We restrict the directions in which the singularities are resolved to ensure we produce not just immersions but embeddings. We assume that in the disk of radius $1/10$ centered at each singularity, both $K_1$ and $K_2$ consist of linear segments intersecting transversely, as shown in Figure ~\ref{fig:disks}. Fix $\ep>0$ so that the intervals $[t_i-\ep,t_i+\ep]$, $i=1,2, \ldots,5$, are disjoint and $K_1([t_i-\ep,t_i+\ep])$ is contained in $B_{\frac{1}{10}}(K_1(s_i))$ for $i=1,2,\ldots,5$. These intervals are the strands we will move to resolve the singularities.
Let $w_1,\ldots, w_5$ be the unit tangent vectors to each line segment at the singular points of $K_1$. Fix $\delta >0$ so that $\{v\in S^{d-2} :\;\parallel v-w_1\parallel<\delta\}$ and $\{v\in S^{d-2} :\;\parallel v-w_4\parallel<\delta\}$ are disjoint. As mentioned above, we avoid moving the third strand off of the triple point in these directions to prevent the introduction of a double point.
\begin{figure}[h]
\begin{center}
$$\includegraphics[width=4in]{delta_disks_1.png}$$
\caption{$B_{\frac{1}{10}}(K_1(t_1))\cap K_1$ and $B_{\frac{1}{10}}(K_1(t_2))\cap K_1$.}
\label{fig:disks}
\end{center}
\end{figure}
We produce a manifold $\mathcal{M}_\beta$ as the image of a topological manifold $M_\beta$ embedded in $\Emb(\hat{S}^1,\R^d)$ by resolving singular knots with triple and double points. The manifold $\mathcal{M}_\beta$ decomposes as the union $\bigcup_{i=1}^6 \mathcal{M}_i$, where each $\mathcal{M}_i$ is the image in $\Emb(\hat{S}^1,\R^d)$ of a resolution map defined below. The domains of the resolution maps for the main pieces, $\mathcal{M}_1$ and $\mathcal{M}_2$, are denoted $M_1$ and $M_2$ and are homeomorphic to $\left( S^{d-2} \setminus \cup_4 B_\delta\right) \times S^{d-3}\times S^{d-3}$. The domains of resolution maps defining the remaining four families are denoted $M_i\times \I$, where $M_i$ is homeomorphic to $S^{d-3}\times S^{d-3}\times S^{d-3}$ for $i=3,4,5,6$.
\begin{definition}
For any triple $(\ep_3,\ep_4,\ep_5)$ with each $\ep_i\leq \ep$ for $\ep$ as above, define \[M_{1}(\ep_3,\ep_4,\ep_5) \subset \Imm_{\geq \beta_{1}}(\hat{S}^1,\R^d) \times \prod_{k=3}^5 (S^{d(r_{k}) }\times \I \times \I)\] as the subspace of all
$K_{1} \times \prod (v_{i}, a_{i}, \varepsilon_i),$ where $a_3 = \frac{1}{10}$, $a_4=a_5 = \frac{\delta}{10}$, and $v_3$ is such that the distances between $v_3$ and the vectors $\pm w_1$ and $\pm w_4$ are all greater than or equal to $\delta$. There are no restrictions on $v_4,v_5\in S^{d-3}$.
\end{definition}
We will suppress the dependence of $M_1$ on the values of $\ep_3,\ep_4,\ep_5 \leq \ep$ as well as $\delta$ except when needed.
\begin{lemma}
The restriction of $\rho_{\beta_1,S}$ to $M_{1}$ maps to $\Emb(\hat{S}^1,\R^d) \subset \Imm_{\geq \phi}(\hat{S}^1,\R^d)$.
\end{lemma}
Choose the immersion $K_2$ as shown in Figure \ref{fig:K1K2}, and assume that the constants $\delta >0$ and $\varepsilon>0$ chosen above satisfy similar conditions for $K_2$, to define $M_2$ analogously. The restriction of $\rho_{\beta_2,S}$ maps $M_2$ to $\Emb(\hat{S}^1,\R^d) \subset \Imm_{\geq \phi}(\hat{S}^1,\R^d)$. We denote the families of embeddings $\rho_{\beta_1,S}(M_1)$ and $\rho_{\beta_2,S}(M_2)$ by $\mathcal{M}_1$ and $\mathcal{M}_2$ respectively, and connect the boundary components of $\mathcal{M}_1$ to those of $\mathcal{M}_2$ to build a family without boundary.
Each boundary component can also be described as the family of knots obtained by resolving a singular knot with three double points. In fact, resolving the triple point in $K_1$ by moving the strand $K_1\left([t_3-\varepsilon_3, t_3+\varepsilon_3]\right)$ in the direction of $\pm w_1$ or $\pm w_4$ yields an immersion with three double points. The four boundary components of $\mathcal{M}_1$ are families of resolutions of these four knots.
\begin{definition} Let $K_3, K_4, K_5$ and $K_6$ be the singular knots, each with three double points, defined below and shown in Figure \ref{fig:K3to6}.
\begin{align*} K_3 & = \rho_{3}\left(K_1, w_4, \tfrac{1}{10}, \ep_3 \right) \\
K_4 & = \rho_{3}\left(K_1, -w_4, \tfrac{1}{10}, \ep_3\right)\\
K_5 & = \rho_{3}\left(K_1, w_1,\tfrac{1}{10}, \ep_3\right) \\
K_6 & = \rho_{3}\left(K_1,-w_1, \tfrac{1}{10}, \ep_3 \right) \end{align*} \end{definition}
\begin{figure}[h]
\begin{center}
$$\includegraphics[width=3in]{K3to6.png}$$
\caption{Singular knots $K_3$, $K_4$, $K_5$, and $K_6$.}
\label{fig:K3to6}
\end{center}
\end{figure}
We resolve these knots, restricting the directions so the resulting embeddings are those in the boundary components of $\mathcal{M}_1$. Initially, we focus on $K_3$. The double points corresponding to $[x_1,x_4]$ and $[x_2,x_6]$, labeled $a$ and $c$, are resolved in the same way as the double points in $K_1$. The double point corresponding to $[x_3,x_5]$, labeled $b$, is resolved using only vectors in the direction $v-w_4$ for some $v$ such that $\parallel v - w_4 \parallel = \delta$. This guarantees that resolving this double point in $K_3$ yields the $\parallel v_3 - w_4\parallel = \delta$ boundary component of $\mathcal{M}_1$.
\begin{definition}
Define $M_3(\ep_3,\ep_4,\ep_5) \subset \Imm_{\geq \beta_3}(\hat{S}^1,\R^d) \times \prod_{i=3,4,6}\left( S^{d-3}\times \I\times \I\right)$ where $\beta_3=[x_1,x_4]\cdot[x_2,x_6]\cdot [x_3,x_5]$ as the subset of all $K_3\times \prod_{i = 3,4,6}\left(u_i,\frac{\delta}{10},\varepsilon_i\right)$ where $u_4$ and $u_6$ are unrestricted and $u_3$ satisfies $\parallel w_4 +\delta u_3\parallel = 1$.
\end{definition}
\begin{proposition}\label{prop:resolboundary}
Let $S_3 = \{x_3,x_4,x_6\}$. The restriction of $\rho_{\beta_3, S_3}$ maps $M_{3}$ to $\Emb(\hat{S}^1,\R^d) \subset \Imm_{\geq \phi}(\hat{S}^1,\R^d)$, and $\rho_{\beta_3,S_3}(M_3)$ is the $\parallel v_3 - w_4\parallel = \delta$ boundary component of $\mathcal{M}_1$.
\end{proposition}
\begin{proof}
The resolution $\rho_{\beta_3,S_3}(K_3) = \rho_{\beta_3}\left( \rho_{3}\left( K_1,w_4,\frac{1}{10},\ep_3\right) \right)$ using $u_3$ as in the definition of $M_3(\ep_3,\ep_4,\ep_5)$ is the same embedding as the resolution $\rho_{\beta_1} (K_1)$ using $v_3 = w_4+\delta u_3$, since \[\frac{1}{10} w_4\,\exp \left(\frac{1}{(t-t_3)^2+\ep_3^2}\right) + \frac{\delta}{10}u_3 \exp\left(\frac{1}{(t-t_3)^2+\ep_3^2}\right) = \frac{1}{10} v_3 \exp \left(\frac{1}{(t-t_3)^2+\ep_3^2}\right).\] \end{proof}
Similarly resolving the knots $K_4, K_5$, and $K_6$ yields the boundary components of $\mathcal{M}_1$ corresponding to $\parallel v_3 + w_4\parallel = \delta$, $\parallel v_3 - w_1\parallel = \delta$, and $\parallel v_3 + w_1\parallel = \delta$ respectively. This process can also be applied to the boundary components of $\mathcal{M}_2$. Let $K_7, K_8, K_9,$ and $K_{10}$ be the four singular knots obtained from $K_2$ by moving $K_2\left([t_3-\ep_3,t_3+\ep_3]\right)$ in the direction of the tangent vectors to the other two strands intersecting at the triple point, as shown in Figure \ref{fig:k7to10}. As with $K_1$, resolving these singular knots gives the four boundary components of $\mathcal{M}_2$.
\begin{figure}[h]
\begin{center}
$$\includegraphics[width=3in]{K7to10.png}$$
\caption{Singular knots $K_7$, $K_8$, $K_9$, and $K_{10}$.}
\label{fig:k7to10}
\end{center}
\end{figure}
Since each of the four knots $K_3,\ldots,K_6$ has the same singularity data as one of $K_7,\ldots,K_{10}$, we have four pairs of knots which are isotopic in $\Imm_{=\beta_i}(\hat{S}^1,\R^4)$, and thus in $\Imm_{=\beta_i}(\hat{S}^1,\R^d)$ with $d\geq 4$, where $\beta_3,\ldots,\beta_6$ each encodes singularity data for a knot with exactly three double points. If $d>4$, we require that the isotopy be through knots in $\R^4\subset \R^d$ (with the standard embedding). If $d = 4$, we restrict the steps of the isotopy, as described in the Appendix, to simplify evaluation of Longoni cocycle on the cycle. Resolving each singular knot in these four isotopies yields four families, denoted $\mathcal{M}_3, \mathcal{M}_4,\mathcal{M}_5,$ and $\mathcal{M}_6$, parametrized by $S^{d-3}\times S^{d-3}\times S^{d-3}\times\I$. Specifically, if $h_i:\I \rightarrow \Imm_{\geq \beta}(\hat{S}^1,\R^d)$ is an isotopy, then these $\mathcal{M}_i$ are be the images of the composites
\begin{center} \begin{multline}\begin{CD} M_i\times \I = S^{d-3}\times S^{d-3}\times S^{d-3}\times\I @>{\mathrm{Id}\times h_i}>> \\ S^{d-3}\times S^{d-3}\times S^{d-3}\times\Imm_{= \beta_i}({S}^1,\R^d) @>{\rho_{\beta_i,S_i}}>> \Emb(\hat{S}^1,\R^d). \end{CD}\end{multline} \end{center}
For $i=3,4,5,6$, the boundary of $\mathcal{M}_i$ is the disjoint union of a boundary component of $\mathcal{M}_1$ and a boundary component of $\mathcal{M}_2$, providing a way to glue the boundary of $\mathcal{M}_1$ to the boundary of $\mathcal{M}_2$.
The union of these six $(3d-8)$-dimensional families in $\Emb(\hat{S}^1,\R^d)$ gives a single family without boundary. Let \[M_\beta = \left(M_1\sqcup M_2 \sqcup\left( \sqcup_{i=3}^6 M_i\times \I\right) \right)/ \sim\] where each boundary component of $M_3,\ldots, M_6$ is identified with a boundary component of $M_1$ or $M_2$ so as to be compatible with Proposition ~\ref{prop:resolboundary}. Let $\mathcal{M}_\beta$ be the image of the orientable topological manifold $M_\beta$ under the resolution map defined above. For $d=4$, the resolution map takes $M_\beta$ to $\Emb(S^1,\R^d)$, as the isotopies we have chosen do not respect the fixed basepoint.
\begin{theorem} If $d>4$ is even then the fundamental class of $\mathcal{M}_\beta$ is a non-trivial homology class in $\Emb(\hat{S}^1,\R^d)$ for any choice of isotopies $h_i$ through $\Imm_{=\beta_i}(\hat{S}^1,\R^d)$. For $d=4$ the fundamental class of $\mathcal{M}_\beta$ is a non-trivial homology class in $\Emb(S^1,\R^d)$ if the isotopies $h_i$ satisfy a sequence of specified steps.
\end{theorem}
For more details on the case $d=4$, see the Appendix and \cite{Pela12}. To prove $[\mathcal{M}_\beta]$ is nontrivial, we evaluate a cocycle due to Longoni \cite{Long04} on $[\mathcal{M}_\beta]$ using configuration space integrals. This is the main result of Section 4.
\section{The Longoni cocycle}
In \cite{CCL02}, Cattaneo, Cotta-Ramusino, and Longoni use configuration space integrals to define a chain map $I$ from a complex of decorated graphs to the de Rham complex of $\Emb({S}^1,\R^d)$. The starting point is the evaluation map $ev: C_q[S^1]\times \Emb(S^1,\R^d)\rightarrow C_q[\R^d]$, where $C_q[M]$ is the Fulton-MacPherson compactified configuration space. See \cite{Sinh04} for more details. For some graphs $G$ (namely those with no internal vertices), the image of the chain map $I$ is defined by pulling back a form determined by $G$ from $C_q[\R^d]$ to $C_q[S^1]\times\Emb(S^1,\R^d)$ and then pushing forward to $\Emb(S^1,\R^d)$.
To understand the general case, let $ev^*C_{q,r}[\R^d]$ be the total space of the pull-back bundle shown below: \[ \xymatrix{ev^*C_{q,r}[\R^d] \ar[r]^{\hat{ev}} \ar[d] & C_{q+r}[\R^d] \ar[d] \\ C_q^{ord}[S^1]\times \Emb({S}^1,\R^d) \ar[r]^{\qquad ev} & C_q[\R^d] \; ,}\] where $C_q^{ord}[S^1]$ is the connected component of $C_q[S^1]$ in which the ordering on the points in the configuration agrees with the ordering induced by the orientation of $S^1$. Fix an antipodally symmetric volume form on $S^{d-1}$, denoted $\alpha$. A choice of $\alpha$ determines tautological $(d-1)-$forms on $ev^*C_{q,r}[\R^d]$, defined by \[ \theta_{ij} = \hat{ev}^* \phi_{ij}^*(\alpha)\] where $\phi_{ij}: C_q(\R^d)\rightarrow S^{d-1}$ sends a configuration to the unit vector from the $i-$th point to the $j-$th point in the configuration. We use integration over the fiber of the bundle $ev^*C_{q,r}[\R^d] \rightarrow \Emb(S^1,\R^d)$, which is the composite of the projections \[ ev^*C_{q,r}[\R^d]\rightarrow C_{q}^{ord}[S^1]\times \Emb({S}^1,\R^d) \rightarrow \Emb({S}^1,\R^d),\] to push forward products of the tautological forms to forms on $\Emb({S}^1,\R^d)$. Which forms to push forward will be determined by graphs.
Consider connected graphs which satisfy the following conditions. A \textit{decorated graph (of even type)} is a connected graph consisting of an oriented circle, vertices on the circle (called \textit{external vertices}), vertices which are not on the circle (called \textit{internal vertices}), and edges. We require that all vertices are at least trivalent. The decoration consists of an enumeration of the edges and an enumeration of the external vertices that is cyclic with respect to the orientation of the circle. We will call the portion of the oriented circle between two external vertices an \textit{arc}.
\begin{definition} Let $\D_e$ be the vector space generated by decorated graphs of even type with the following relations. We set $G = 0$ if there are two edges in $G$ with the same endpoints, or if there is an edge in $G$ whose endpoints are the same internal vertex. The graphs $G$ and $G'$ are equal if they are isomorphic as graphs and the enumerations of their edges differ by an even permutation.
\end{definition}
The vector space $\D_e$ admits a bigrading as follows. Let $v_e$ and $v_i$ be the number of external and internal vertices, respectively, and let $e$ be the number edges. The order of a graph is given by \[\mathrm{ord}\, G = e - v_i\] and the degree of a graph is defined by \[\deg G = 2e - 3v_i -v_e.\] Let $\mathcal{D}_e^{k,m}$ be the vector space of equivalence classes with order $k$ and degree $m$. In \cite{CCL02}, Cattaneo, Cotta-Ramusino and Longoni define a map from this vector space to the space of ${(m+(d-3)k)-}$ forms on $\Emb(S^1, \R^d)$.
\begin{definition} Define $I(\alpha): \mathcal{D}_e^{k,m} \rightarrow \Omega^{m+(d-3)k} \left(\Emb({S}^1,\R^d)\right)$ as follows. \begin{enumerate}
\item Choose an ordering on the internal vertices.
\item Associate each edge in $G$ joining vertex $i$ and vertex $j$ to the tautological form $\theta_{ij}$.
\item Take the product of these tautological forms with the order of multiplication determined by the enumeration of the edges, to define a form on $ev^*C_{q,r}[\R^d]$.
\item Integrate this form over the fiber to obtain a form on $\Emb(S^1,\R^d)$. \end{enumerate}
\end{definition}
This integration over the fiber defines the pushforward and in this case is often called a configuration space integral. There is a coboundary map on $\D_e$ which makes $I(\alpha)$ a cochain map.
\begin{definition} Define a coboundary operator on $\D_e$ by taking $\delta G$ to be the signed sum of the decorated graphs obtained from $G$ by contracting, one at a time, the arcs of $G$ and the edges of $G$ which have at least one endpoint at an external vertex. After contracting, the edges and vertices are relabeled in the obvious way - if the edge (respectively vertex) labeled $i$ is removed, we replace the label $j$ by $j-1$ for all $j>i$. When contracting an arc joining vertex $i$ to $i+1$, the sign is given by $\sigma(i,i+1) = (-1)^{i+1}$, and when contracting the arc joining vertex $j$ to vertex $1$, the sign is given by $\sigma(j,1) = (-1)^{j+1}$. When contracting the edge $l$, the sign is given by $\sigma(l) = l + 1 + v_e$, where $ v_e$ is the number of external vertices.
\end{definition}
\begin{theorem} \cite{CCL02}
The map $I(\alpha)$ determines a cochain map and therefore induces a map on cohomology, which we denote $I(\alpha): H^{k,m}(\mathcal{D}_e)\rightarrow H^{m+(d-3)k}(\Emb(\hat{S}^1,\R^d))$.\end{theorem}
At the level of forms, $I(\alpha)$ depends on the choice of antipodally symmetric volume form $\alpha$. On cohomology, when $d > 4$ this is independent of $\alpha$.
\begin{example} From \cite{CCL02}, we have the graph cocycle shown in Figure \ref{fig:ccl_cocycle}, originally investigated by Bott and Taubes \cite{BoTa94} for $d=3$.
\begin{figure}[h]
\begin{center}
$$\includegraphics[width=2.25in]{ccl_cocycle.png}$$
\caption{Graph cocycle given by Cattaneo et al. in \cite{CCL02}.}
\label{fig:ccl_cocycle}
\end{center}
\end{figure}
This induces the cocycle \[ \frac14 \int_{ev^*C_{4,0}[\R^d]}\theta_{13}\theta_{24} - \frac13\int_{ev^*C_{3,1}[\R^d]}\theta_{14}\theta_{24}\theta_{34} \in H^{2d - 6}\left(\Emb(\hat{S}^1,\R^d)\right).\]
\end{example}
In \cite{CCL02}, Cattaneo et al. show that this cocycle evaluates non-trivially on $\rho_{[x_1,x_3]\cdot[x_2,x_4]}\left(K\times S^{d-3}\times S^{d-3}\right),$ where $K$ is a singular knot with two double points respecting $[x_1,x_3]\cdot[x_2,x_4]$ (in this case, the cycle does not depend on the ordered subset $S\subseteq \{x_1,x_2,x_3,x_4\}$).
\begin{example} In \cite{Long04}, Longoni gives the example shown in Figure \ref{fig:longoni_cocycle} of a graph cocycle $G_L$ in $H^{3,1}(\D_e)$ which uses nontrivalent graphs. There $I(\alpha)\left(G_L\right)\in H^{3(d-3)+1}(\Emb(\hat{S}^1,\R^d))$ is the form \[\omega = \int_{ev^*C_{4,1}[\R^d]} \theta_{15}\theta_{45}\theta_{35}\theta_{25} + 2\int_{ev^*C_{5,0}[\R^d]} \theta_{13}\theta_{14}\theta_{25}.\] We pair this cocycle with the cycle $[M_\beta]$ defined in Section ~2 to see that both are nontrivial.
\begin{figure}[h]
\begin{center}
$$\includegraphics[width=2.5in]{longoni_cocycle.png}$$
\caption{Graph cocycle given by Longoni in \cite{Long04}.}
\label{fig:longoni_cocycle}
\end{center}
\end{figure}
\end{example}
\section{Nontriviality}
\begin{proposition} Assume $d> 4$ is even. Let $[\mathcal{M}_\beta]\in H_{3(d-3)+1}(\Emb(\hat{S}^1,\R^d))$ be the cycle defined in Section ~\ref{sec:cycledef}, and let $\omega\in H^{3(d-3)+1}(\Emb(\hat{S}^1,\R^d))$ be the Longoni cocycle defined in the last section. Then $\omega([\mathcal{M}_\beta]) = \pm 2$. In particular, $\omega([\mathcal{M}_\beta])$ is nonzero, and therefore both $\omega$ and $[\mathcal{M}_\beta]$ are non-trivial.
\end{proposition}
In \cite{Tour07}, Turchin calculates that $E^2_{-5,3(d-1)}$ has rank one, so $[\mathcal{M}_\beta]$ is a generator of this group. The proposition also holds for $d=4$ if $\Emb(\hat{S}^1,\R^d)$ is replaced by $\Emb(S^1,\R^d)$.
\begin{proof}
First we show that $\omega_2([\mathcal{M}_\beta])=\pm 1$. Let $g: ev^*C_{5,0}[\R^d]\rightarrow S^{d-1}\times S^{d-1}\times S^{d-1}$ be the map shown in the diagram below, where $\bar{\psi}=\phi_{13}\times \phi_{14}\times \phi_{25}$. Then $\omega_2$ is the pushforward along $\pi:ev^*C_{5,0}[\R^d]\rightarrow \Emb(\hat{S}^1,\R^d)$ of $g^*(\alpha\otimes\alpha\otimes\alpha)$.
\[\xymatrix{ &ev^*C_{5,0}[\R^d]\ar[r] \ar[d] \ar@/^2pc/[rr]^g\ar@/_6pc/[dd]_\pi & C_5[\R^d] \ar[r]^{\bar{\psi}\qquad \qquad}\ar[d]^{id} & S^{d-1}\times S^{d-1}\times S^{d-1} \\ & C_5^{ord}[S^1]\times \Emb(\hat{S}^1,\R^d)\ar[r]\ar[d] & C_5[\R^d] \\ \mathcal{M}_\beta \ar@^{(->}[r] & \Emb(\hat{S}^1,\R^d) }\]
By naturality of pushforwards, $\omega_2([\mathcal{M}_\beta])=g^*(\alpha\otimes\alpha\otimes\alpha)([\pi^{-1}(\mathcal{M}_\beta)])$. The bundle $\pi: ev^*C_{5,0}[\R^d]\rightarrow \Emb(\hat{S}^1,\R^d)$ is trivial, so $g^*(\alpha\otimes\alpha\otimes\alpha)([\pi^{-1}(\mathcal{M}_\beta)])= \int_{C_5^{ord}[S^1]\times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes\alpha)$.
To calculate $\int_{C_5^{ord}[S^1]\times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes\alpha)$, we first partition $C_5^{ord}[S^1]$. For $i=1,\ldots, 5$ let $N_i=(t_i-\varepsilon, t_i+\varepsilon)$, where the $t_i$ are the times of singularity in $K_1$ and $K_2$, and $\varepsilon$ is as in Section ~\ref{sec:cycledef}. Define \[ C_5^{(i)} = \left\{ \bar{s}\in C_5^{ord} : s_j\not\in N_i\; \mathrm{for}\; j=1,\ldots, 5 \; \mathrm{and}\; \bar{s}\notin C_5^{(m)} \; \mathrm{for} \;m<i\right\},\] and $C_5^c = C_5^{ord}[S^1]\backslash\left( \cup_{i=1}^5 C_5^{(i)}\right)$, so $C_5^c$ is the set of all $\bar{s}\in C_5^{ord}[S^1]$ such that $t_i-\varepsilon < s_i< t_i+\varepsilon$ for $i=1,\ldots, 5$. Then $C_5^{ord}[S^1]$ decomposes as $C_5^{ord}[S^1]= C_5^c\sqcup C_5^{(1)}\sqcup \cdots \sqcup C_5^{(5)}$, and we obtain a corresponding decomposition of $\int_{C_5^{ord}[S^1]\times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes \alpha)$. We will show that $\int_{C_5^{(m)}\times \mathcal{M}_\beta}g^*(\alpha\otimes\alpha\otimes\alpha) = 0$ for $m=1,\ldots, 5$, so calculating $\omega_2([\mathcal{M}_\beta])$ reduces to evaluating the integrals \[\int_{C_5^c\times \mathcal{M}_i} g^*(\alpha\otimes\alpha\otimes\alpha).\]
For $m=3,4,5$, we show $\int_{C_5^{(m)}\times \mathcal{M}_\beta}g^*(\alpha\otimes\alpha\otimes\alpha) = 0$ by showing $\int_{C_5^{(m)}\times \mathcal{M}_i}g^*(\alpha\otimes\alpha\otimes\alpha) = 0$ for $i = 1,\ldots,6$. Recall that manifolds have only trivial forms in degrees above their dimension, so a form pulled back through a smaller dimensional manifold is always zero. To prove that the integrals $\int_{C_5^{(m)}\times \mathcal{M}_i}g^*(\alpha\otimes\alpha\otimes\alpha) $ are zero, we show that the map $g$ factors through spaces of smaller dimension when restricted to each of the subspaces $C_5^{(m)}\times \mathcal{M}_i$.
First, consider the case $\int_{C_5^{(3)}\times \mathcal{M}_1} g^*(\alpha\otimes\alpha\otimes\alpha)$. Recall that $\mathcal{M}_1$ is $\rho_{\beta_1,S}\left(K_1\times \prod_{k=3}^5 (v_k,a_k,\varepsilon)\right) $. If $t\notin N_3$ and $\gamma \in \mathcal{M}_1$, the point $\gamma(t)$ does not depend on the value of $v_3$ in the preimage of $\gamma$. This gives us the following factorization of $g\big|_{C_5^{(3)}\times \mathcal{M}_1}$:
\[ \xymatrix{C_5^{(3)}\times \mathcal{M}_1 \ar[rr]^g \ar[dr] & & S^{d-1}\times S^{d-1}\times S^{d-1} \\ & C_5^{(3)}\times S^{d-3}\times S^{d-3} \ar[ur] & } \]
Since $\dim(C_5^{(3)}\times S^{d-3}\times S^{d-3} ) = 2d-1$ is less than $\dim(S^{d-1}\times S^{d-1}\times S^{d-1}) = 3d-3$, we have $\int_{C_5^{(3)}\times \mathcal{M}_1}g^*(\alpha\otimes\alpha\otimes\alpha)=0$.
Similarly, for $m=4$ or $m=5$, the restriction $g\big|_{C_5^{(m)}\times\mathcal{M}_1}$ factors through \[C_5^{(m)} \times \{v_3\in S^{d-2}\,:\, \parallel v_3\pm w_1 \parallel > \delta\; \mathrm{and} \parallel v_3\pm w_4\parallel >\delta\}\times S^{d-3},\] so the corresponding integrals are zero. This argument also shows that $\int_{C_5^{(m)}\times \mathcal{M}_2} g^*(\alpha\otimes\alpha\otimes\alpha) = 0$ for $m=3,4,5$. For $i=3,4,5,6$ and $m=3,4, 5$, the restriction $g\big|_{C_5^{(m)}\times \mathcal{M}_i}$ factors through $S^{d-3}\times S^{d-3}\times \I$ and therefore $\int_{C_5^{(m)}\times \mathcal{M}_i}g^*(\alpha\otimes\alpha\otimes\alpha)$ is zero.
We show $\int_{C_5^{(1)}\times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes\alpha) = 0$ by replacing $\mathcal{M}_\beta$ with the family of embeddings obtained by moving the first strand (instead of the fourth) off of the double point $K_i(t_1) =K_i(t_4)$, over which $g^*$ factors through a space of lower dimension. We replace $\mathcal{M}_\beta$ in two steps - first with the family of embeddings in which both strands are moved off the double point, and then by the family in which only the first strand is moved.
Let $\mathcal{M}_\beta'$ be the piecewise smooth subspace of $\Emb(\hat{S^1}, \R^d)$ defined similarly to $\mathcal{M}_\beta$, but by choosing the ordered subset of variables in $\beta_1$ and $\beta_2$ to be $S=\{x_1,x_3,x_4,x_5\}$, and fixing $a_1=a_4$ and $v_1=-v_4$. In other words, $\mathcal{M}_\beta'$ is obtained from $K_1,\ldots, K_6$ by moving both strands off the double point $K_i(t_1)=K_i(t_4)$ in antipodal directions.
We define a cobordism $W_1$ between $\mathcal{M}_\beta$ and $\mathcal{M}_\beta'$ as the subspace of $\Emb(\hat{S^1}, \R^d)$ parametrized by $\left(\sqcup_i M_i\right)\times \I$, with the embedding corresponding to the parameter $u\in \I$ determined by $a_1 = u a_4$ (so the $\I$ parametrizes how far the strand with $K_i(t_1)$ is moved off the double point).
By Stokes' theorem, \[ \int_{C_5^{(1)}\times W_1} dg^*(\alpha\otimes\alpha\otimes\alpha) \\ = \int_{\partial(C_5^{(1)}\times W_1)} g^*(\alpha\otimes\alpha\otimes\alpha).\] Since $dg^*(\alpha\otimes\alpha\otimes\alpha) =g^*d(\alpha\otimes\alpha\otimes\alpha) = 0$, we have \begin{equation}\label{stokes} 0 = \int_{\partial C_5^{(1)}\times W_1} g^*(\alpha\otimes\alpha\otimes\alpha) +\int_{C_5^{(1)}\times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes\alpha)-\int_{C_5^{(1)}\times \mathcal{M}_\beta'} g^*(\alpha\otimes\alpha\otimes\alpha).\end{equation} The restriction $g^*\big|_{\partial C_5^{(1)}\times W_1}$ factors through $\partial C_5^{(1)}\times \left(\sqcup_i \mathcal{M}_i\right)$. If $\bar{s}\in \partial C_5^{(1)}$ then the parameter, $u\in \I$ determining how far the first strand is moved does not affect $g(\bar{s}, \gamma)$ for $\gamma\in W_1$. Thus, $\int_{\partial C_5^{(1)}\times W_1} g^*(\alpha\otimes\alpha\otimes\alpha) = 0$ and \[ \int_{C_5^{(1)}\times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes\alpha)=\int_{C_5^{(1)}\times \mathcal{M}_\beta'} g^*(\alpha\otimes\alpha\otimes\alpha).\]
Let $\mathcal{M}_\beta''$ be the piecewise smooth subspace of $\Emb(\hat{S^1}, \R^d)$ obtained by choosing the ordered subset of variables in $\beta_1$ and $\beta_2$ to be $S=\{x_1,x_3,x_5\}$. In other words, $\mathcal{M}_\beta''$ is obtained from $K_1,\ldots, K_6$ by moving only the first strand off the double point $K_i(t_1)=K_i(t_4)$. Let $W_2\subset \Emb(\hat{S^1}, \R^d)$ be parametrized by $\left(\sqcup_i M_i\right)\times \I$, with the embedding corresponding to the parameter $u\in \I$ given by choosing $a_4'' = u a_4$ (so the interval parametrizes how far the strand with $K_i(t_4)$ is moved off the double point). Then $W_2$ gives a cobordism between $\mathcal{M}_\beta'$ and $\mathcal{M}_\beta''$, as $\partial W_2 = \mathcal{M}_\beta \sqcup (-\mathcal{M}_\beta')$. Using Stokes' Theorem and naturality again, we have \begin{equation}\label{stokes2} 0 = \int_{\partial C_5^{(1)}\times W_2} g^*(\alpha\otimes\alpha\otimes\alpha) +\int_{C_5^{(1)}\times \mathcal{M}_\beta'} g^*(\alpha\otimes\alpha\otimes\alpha)-\int_{C_5^{(1)}\times \mathcal{M}_\beta''} g^*(\alpha\otimes\alpha\otimes\alpha).\end{equation}
The restriction $g^*\big|_{\partial C_5^{(1)}\times W_2}$ does not factor through $\partial C_5^{(1)}\times \left(\sqcup_i M_i\right)$. To show the first integral in \eqref{stokes2} is zero, we consider $W_2$ as a subspace of $\Imm_{\leq [x_1,x_2], t_1,t_4}(\hat{S}^1,\R^d)$, the subset of $\Imm(\hat{S}^1,\R^d)$ consisting of all immersions $\gamma$ with at most one singularity - a double point $\gamma(t_1)=\gamma(t_4)$. Since a configuration in $\partial C_5^{(1)}$ does not contain the point $t_1$, the map $g$ is well-defined on $\partial C_5^{(1)}\times \mathrm{Imm}_{\leq [x_1,x_2],t_1,t_4}(\hat{S}^1,\R^d)$. Letting the dependance on the lengths of the strands be apparent, we now work with $W_2=W_2(\ep_3,\ep_4,\ep_4)$ as a subspace of $\mathrm{Imm}_{\leq [x_1,x_2],t_1,t_2}(\hat{S}^1,\R^d)$. In this larger space, $W_2(\ep_3,\ep_4,\ep_5)$ is cobordant to $W_2(\ep_3,0,\ep_5)$. The cobordism is given by $W_3\subset \Imm_{[x_1,x_2],t_1,t_4}(\hat{S}^1,\R^d)$ parametrized by $\left(\sqcup_i M_i\right)\times \I\times \I$ where the second unit interval parametrizes the length of the strand centered at $t_4$ moved by the resolution map.
By Stokes' Theorem and naturality, \[ 0=\int_{\partial C_5^{(1)} \times W_3} dg^*(\alpha\otimes\alpha\otimes\alpha) = \int_{\partial\left(\partial C_5^{(1)} \times W_3\right)}g^*(\alpha\otimes\alpha\otimes\alpha), \] and thus, \begin{multline} 0 = \int_{\partial (\partial C_5^{(1)})\times W_3}g^*(\alpha\otimes\alpha\otimes\alpha) + \int_{\partial C_5^{(1)}\times W_2(\ep_3,\ep_4,\ep_5)} g^*(\alpha\otimes\alpha\otimes\alpha) - \int_{\partial C_5^{(1)}\times W_2(\ep_3,0,\ep_5)} g^*(\alpha\otimes\alpha\otimes\alpha)\\ = \int_{\partial C_5^{(1)}\times W_2(\ep_3,\ep_4,\ep_5)} g^*(\alpha\otimes\alpha\otimes\alpha). \end{multline} The second equality holds because $\partial(\partial C_5^{(1)} )= \varnothing$ and the dimension of $W_2(\ep_3,0,\ep_5)$ is $2d-3$.
By the same argument, $\int_{C_5^{(5)}\times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes\alpha)=0$. Calculating $\int_{C_5\times \mathcal{M}_\beta }g^*(\alpha\otimes\alpha\otimes\alpha)$ thus reduces to calculating $\int_{C_5^c\times \mathcal{M}_i} g^*(\alpha\otimes\alpha\otimes\alpha)$ for $i=1,\ldots, 6$.
We chose the antipodally symmetric volume form, $\alpha$, to be concentrated near the points $\bar x_1 = (0,\ldots,0,1)\in S^{d-1}$ and $\bar x_2 = (0,\ldots,0,-1)\in S^{d-1}$. Let $\tau_{\bar x_1}$ and $\tau_{\bar x_2}$ be the Thom classes of these points, as defined in Section 6 of \cite{BoTu82}, so $\alpha = \frac{1}{2}\left(\tau_{\bar x_1} + \tau_{\bar x_2}\right)$. Let $y$ be the arc in $S^{d-1}$ connecting $(0,\ldots,0,1)$ and $(0,\ldots,0,-1)$, defined as \[y = \left\{ \left(0,\ldots,0,\sqrt{1-s^2},s\right)\in S^{d-1} \; : \; -1\leq s\leq 1\right\}.\] The Thom class $\tau_y$ of $y$ can be chosen so that $d\tau_y = \tau_{\bar x_1} - \tau_{\bar x_2} = 2(\tau_{\bar x_1} - \alpha)$.
We have \begin{align*} \int_{C_5^c\times \mathcal{M}_i} g^*(\alpha\otimes\alpha\otimes\alpha-\tau_{\bar x_1}\otimes \alpha \otimes \alpha) &= \int_{C_5^c\times \mathcal{M}_i} g^*\left( -\tfrac{1}{2} d\tau_y\otimes \alpha \otimes \alpha\right) \\ & = -\tfrac{1}{2} \int_{C_5^c\times \mathcal{M}_i} dg^*(\tau_y\otimes \alpha\otimes \alpha) \\ & =-\tfrac{1}{2}\int_{\partial (C_5^c\times \mathcal{M}_i)} g^*(\tau_y\otimes \alpha\otimes \alpha). \end{align*}
If $(v_1,v_2,v_3) \in g\left(\partial (C_5^c\times \mathcal{M}_i)\right)$ at least one of the first two coordinates of $v_1$ is non-zero, but every $\bar x\in y\subset S^{d-1}$ has $x_1,x_2 = 0$. Thus, the sets $y$ and $g\left(\partial (C_5^c\times \mathcal{M}_i)\right)$ are disjoint and $\int_{C_5^c\times \mathcal{M}_i} g^*(\tau_y\otimes\alpha\otimes\alpha)=0$, which means $\int_{C_5^c\times \mathcal{M}_i} g^*(\alpha\otimes\alpha\otimes\alpha) = \int_{C_5^c\times \mathcal{M}_i} g^*(\tau_{\bar x_1}\otimes \alpha \otimes \alpha)$. By a similar argument, \[\int_{C_5^c\times \mathcal{M}_i} g^*(\alpha\otimes\alpha\otimes\alpha) = \int_{C_5^c\times \mathcal{M}_i} g^*(\tau_{\bar x_1}\otimes \tau_{\bar x_1} \otimes \tau_{\bar x_1}).\] This integral can be calculated by counting the transverse intersections of $g(C_5^c\times \mathcal{M}_i)$ and $(\bar x_1, \bar x_1, \bar x_1)$ in $S^{d-1}\times S^{d-1}\times S^{d-1}$.
Recall that \[g(\bar{s},\gamma)=\left( \frac{\gamma(s_3)-\gamma(s_1)}{\|\gamma(s_3)- \gamma(s_1)\|}, \frac{\gamma(s_4)-\gamma(s_1)}{\|\gamma(s_4)- \gamma(s_1)\|}, \frac{\gamma(s_5)-\gamma(s_2)}{\|\gamma(s_5)- \gamma(s_2)\|}\right).\] Thus, we are counting the number of pairs $(\bar{s},\gamma)\in C_5^c\times \mathcal{M}_i$ for which \[ \frac{\gamma(s_3)-\gamma(s_1)}{\|\gamma(s_3)- \gamma(s_1)\|} = \frac{\gamma(s_4)-\gamma(s_1)}{\|\gamma(s_4)- \gamma(s_1)\|} = \frac{\gamma(s_5)-\gamma(s_2)}{\|\gamma(s_5)- \gamma(s_2)\|} = (0,\ldots,0,1).\] For $\gamma \in\ M_\beta$, this is only possible if $s_i=t_i$ for $i=1,\ldots,5$.
If $\gamma \in \mathcal{M}_1$, then \[ \frac{\gamma(t_3)-\gamma(t_1)}{\|\gamma(t_3)- \gamma(t_1)\|} = \frac{\gamma(t_4)-\gamma(t_1)}{\|\gamma(t_4)- \gamma(t_1)\|} = \frac{\gamma(t_5)-\gamma(t_2)}{\|\gamma(t_5)- \gamma(t_2)\|} = (0,\ldots,0,1)\] exactly when $v_3=v_4=v_5 = (0,\ldots,0,1)$ and so $\int_{C_5^c\times \mathcal{M}_1} g^*(\alpha\otimes\alpha\otimes\alpha) = \pm1$. If $\gamma \in \mathcal{M}_i$ for $i=2,\ldots, 6$, then \[\frac{\gamma(t_3)-\gamma(t_1)}{\|\gamma(t_3)- \gamma(t_1)\|} \neq (0,\ldots, 0, 1) ,\] and $\int_{C_5^c\times \mathcal{M}_i} g^*(\alpha\otimes\alpha\otimes\alpha) = 0$. Thus, $\omega_2([\mathcal{M}_\beta]) = \pm 1$.
Next, we show that $\omega_1([\mathcal{M}_\beta])=0$. Let $f: ev^*C_{4,1}(\R^d)\rightarrow S^{d-1}\times S^{d-1}\times S^{d-1}\times S^{d-1}$ be the map shown in the diagram below, where $\bar \varphi = \phi_{15}\times \phi_{45}\times \phi_{35}\times \phi_{25}$. Then $\omega_1$ is the pushforward of $f^*(\alpha\otimes \alpha\otimes \alpha\otimes\alpha)$ along $p: ev^*C_{4,1}(\R^d)\rightarrow \Emb(\hat{S}^1,\R^d)$.
\[\xymatrix{ &ev^*C_{4,1}(\R^d)\ar[r] \ar[d]^{p_1} \ar@/^2pc/[rr]^f\ar@/_6pc/[dd]_p&C_5[\R^d] \ar[r]^{\bar{\varphi}\qquad \qquad}\ar[d] & S^{d-1}\times S^{d-1}\times S^{d-1}\times S^{d-1} \\ & C_4^{ord}[S^1]\times \Emb(\hat{S}^1,\R^d)\ar[r]\ar[d]^{p_2} & C_4[\R^d] \\ M_\beta \ar@^{(->}[r] & \Emb(\hat{S}^1,\R^d) }\]
Since $p^{-1}(\mathcal{M}_\beta) = p_1^{-1}(C_4^{ord}[S^1]\times \mathcal{M}_\beta)$, we have $\omega_1([\mathcal{M}_\beta])=\int_{p_1^{-1}(C_4^{ord}[S^1]\times \mathcal{M}_\beta)}f^*(\alpha\otimes\alpha\otimes\alpha\otimes\alpha)$. Following the calculation of $\omega_1([\mathcal{M}_\beta])$, define \[ C_4^{(i)} = \left\{ \bar{s}\in C_4^{ord}[S^1] : s_j\not\in N_i\; \mathrm{for}\; j=1,\ldots, 4 \; \mathrm{and}\; \bar{s}\notin C_4^{(m)} \; \mathrm{for} \;m<i\right\}.\] Each configuration in $C_4^{ord}[S^1]$ has four points, so $C_4^{ord}[S^1] = C_4^{(1)}\sqcup\cdots\sqcup C_4^{(5)}$. The arguments used to prove that $\int_{C_5^{(m)} \times \mathcal{M}_\beta} g^*(\alpha\otimes\alpha\otimes\alpha) = 0$ also show $\int_{p_1^{-1}(C_4^{(m)}\times \mathcal{M}_\beta)}f^*(\alpha\otimes\alpha\otimes\alpha\otimes\alpha) = 0$ for $m = 1,\ldots,5$.
\end{proof}
\section{Future Work}
The resolution map in Definition ~\ref{def:resmap} can be generalized to define a resolution map for knots respecting any bracket expression. Instead of choosing an ordered subset of the variables, we repeatedly choose the strands to move so as to resolve the singularity data for the brackets which are not contained inside of any other brackets.
For example, if $(K, \bar{t})$ respects $[[x_1,x_3],[[x_2,x_4],x_5]]$ the point $K(t_1)=K(t_3)$ is first moved away from the point $K(t_2)=K(t_4)=K(t_5)$, turning the original singularity into a double point and a triple point. The double point is then resolved as before, and the triple point is resolved by first moving the fifth strand off the singularity and then resolving the remaining double point.
The description in Section ~\ref{sec:cycledef} of the first differential of the embedding calculus homology spectral sequence is given in terms of ``doubling'' the point $x_i$. In \cite{Pela12} we develop another description of this differential, call it $\tilde{d}_1$, which encodes the singularity data that occurs when a knot respecting a bracket expression is resolved as prescribed in the generalization of the resolution map, but with the directions chosen in such a way as to introduce a new singularity. The boundary components of the family of resolutions of a knot $(K,\bar{t})$ respecting a bracket expression under the generalized resolution map are the same as the families of resolutions of knots respecting the terms in $\tilde{d}_1$ of that bracket expression (with appropriate choices).
Suppose $\beta = \sum_{i=1}^m \beta_i$ is a cycle on the first page of the spectral sequence (where each $\beta_i$ is a bracket expression with a single term) in which the Jacobi identity is not used to simplify the differential. Knots $(K_i, \bar{t})$ respecting the $\beta_i$ can be chosen so that the boundaries of the families of resolutions under the generalized resolution map can be connected by families of embeddings given by an isotopy of underlying singular knots, as in the cycle $[\mathcal{M}_\beta]$ defined here. Thus the process used in this paper can be generalized to more cycles on the first page of the spectral sequence. Because Turchin proved linear duality of the Cattaneo, Cotta-Ramusino, and Longoni graph complex and the $E_1$ page of the embedding calculus spectral sequence, we also have configuration space integrals to evaluate on the families we produce. Together these could give not only a second proof of the collapse of the spectral sequence (Lambrechts, Turchin and Volic use closely related configuration space integrals in their proof of the collapse in \cite{LTV06}), but also geometric representatives and a clear starting point for considering any torsion phenomena.
|
\section{Introduction}
\label{sec:introduction}
Understanding the fundamental causes of the dramatic slowdown of
dynamics when a liquid transforms into a glass is still a subject of
great debate.\cite{Stillinger01,Berthier05,Biroli09,Chandler09}
Essentially all discussion of the glass transition has focused on the
strictly classical regime of liquid state behavior, namely where the
de Broglie wave length is significantly smaller than the particle
size. Given that nearly all known glass forming liquids fall well
within this regime,\cite{Angell96} it is clear that the classical
approximation is generally justified. However there are several
interesting and important examples where quantum fluctuations and
glassiness coexist.\cite{Wu91,Imry09} In such cases, which range from
the behavior of superfluid helium under high pressure to the phase
diagram of quantum random optimization problems, the interplay between
quantum mechanics and the otherwise classical fluctuations that lead
to vitrification can be expected to produce qualitatively novel
physical behavior.\cite{Markland11}
The theoretical investigation of quantum glasses has increased in
recent years. Studies ranging from the investigation of quantum
effects in so-called stripe glasses,\cite{Wolynes03} quantum
spin-glasses~\cite{Cugliandolo98,Cugliandolo99,Cugliandolo01,Biroli01,Cugliandolo02,Cugliandolo04,Cugliandolo06}
and lattice models that mimic the properties of superfluid and
supersolid helium~\cite{Zamponi11} have been presented. In this work
we instead focus on "realistic" off-lattice quantum fluids. The
microscopic detail of our study necessitates the use of
approximations, such as mode-coupling theory (MCT)~\cite{Gotze09} and
ring-polymer molecular dynamics (RPMD),\cite{Manolopoulos04} that are
less well-justified then the methods employed in the studies of the
model systems mentioned above. On the other hand, the approaches used
here have lead to a host of non-trivial predictions both for classical
glass-forming liquids~\cite{Gotze09} as well as a variety of quantum
liquid-state phenomena.\cite{Rabanireview05} We thus expect that the
predictions made in this work to be at least of qualitative accuracy.
The work presented here builds on our earlier report of several novel
effects that arise when glassy dynamics occurs in the quantum
regime.\cite{Markland11} In particular, both RPMD and the quantum
version of mode-coupling theory (QMCT) indicate that the dynamical
phase diagram of glassy quantum fluids is reentrant. As a
consequence, hard-sphere quantum liquids may be forced deeper into the
glass "phase" at fixed volume fraction as quantum fluctuations
increase. This counterintuitive finding has implications not only for
liquid-state systems such as superfluid helium under pressure, but for
a broad class of quantum optimization problems as well.
In comparison to our earlier paper,\cite{Markland11} the work
presented here provides complete details for both the QMCT and the
quantum integral equations needed for generating the required
structural input. In addition, we give a far more extensive
interpretation of the results, largely afforded by our RPMD
simulations. Lastly, we discuss in greater detail the connection of
our results to related theoretical work.
The paper is organized as follows: In Sec.~\ref{sec:qmct} we provide
the details of the QMCT, including a description of the equations for
the density correlator and the mode coupling approximations. In
addition, we discuss the high and low temperature limit of the QMCT
and derive equations for the nonergodic parameter used to determine
the liquid-glass line. In Sec.~\ref{sec:qiet} we describe the quantum
integral equation theory used to obtain the the static input required
by QMCT. Sec.~\ref{sec:rpmd} is devoted to the RPMD method. Results
and discussions are presented in Sec.~\ref{sec:res}. Finally, in
Sec.~\ref{sec:con} we conclude.
\section{A self-consistent quantum mode-coupling theory}
\label{sec:qmct}
The general quantity of interest is the Kubo transform~\cite{Kubo95}
of the time correlation of the collective density operator,
$\hat{\rho}_{\bf q}=\sum_{\alpha=1}^{N} \mbox{e}^{i {\bf q} \cdot {\bf
\hat{r}}_{\alpha}}$, given by
\begin{eqnarray}
\phi_{q}(t) & = & \frac{1}{N\hbar\beta}\int_{0}^{\hbar\beta}
d\lambda\langle \hat{\rho}_{q}^{\dagger}(t+i\lambda)
\hat{\rho}_{q}(0)\rangle\nonumber \\ & \equiv & \frac{1}{N}
\left(\hat{\rho}_{q}(t)| \hat{\rho}_{q}(0)\right),
\label{eq:phiqt-def}
\end{eqnarray}
with a time evolution described by the {\em exact} quantum generalized
Langevin equation (QGLE)~\cite{Rabanireview05}
\begin{equation}
\ddot{\phi}_{q}(t) + \Omega_{q}^{2}\phi_{q}(t) + \int_{0}^{t}d\tau
M_{q}(\tau)\dot{\phi}_{q}(t-\tau) = 0,
\label{eq:qgle}
\end{equation}
In the above, we have used the notion that ${\bf \hat{r}}_{\alpha}$
stands for the position vector operator of particle $\alpha$ with a
conjugate momentum ${\bf \hat{p}}_{\alpha}$ and mass $m$, $N$ is the
total number of particles, $\beta=\frac{1}{k_{\mbox{\tiny{B}}}T}$ is
the inverse temperature and $\langle \cdots \rangle$ in
Eq.~(\ref{eq:phiqt-def}) denotes a quantum mechanical ensemble
average. The frequency and memory terms are given by:
\begin{equation}
\Omega_{q}^{2}=\frac{q^{2}}{m \beta\phi_{q}(0)}
\label{eq:omega}
\end{equation}
and
\begin{equation}
M_{q}(t) = \frac{\left(Q_1{\mathcal{L}^{2}}\hat{\rho}_{q} |
\mbox{e}^{i\bar{\mathcal{L}}t}|Q_1{\mathcal{L}}^{2}
\hat{\rho}_{q}\right)} {\Omega_{q}^{2}\phi_{q}(0)},
\label{eq:memory}
\end{equation}
respectively, with ${\mathcal L}=\frac{1}{\hbar}[{\hat H},\cdots]$
being the Liouvillian and $\bar{\mathcal L}=Q_{2}Q_1{\mathcal{L}}Q_1
Q_{2}$. To derive the above equations we have defined two projection
operators (first and second order,
respectively)\cite{Gotze76a,Pathak86}
\begin{equation}
P_1=\left.|\hat{\rho}_{q}\right) \phi_{q}^{-1}(0) \left(\hat{\rho}_{q}|
\right.
\label{eq:projection}
\end{equation}
and
\begin{equation}
P_{2}= \left.|Q_1{\mathcal L} \hat{\rho}_{q} \right) \left(Q_1{\mathcal L}
\hat{\rho}_{q} |Q_1{\mathcal L}\hat{\rho}_{q} \right)^{-1}
\left(Q_1{\mathcal L}\hat{\rho}_{q}|\right.
\label{eq:projection2}
\end{equation}
with $Q_1=1-P_1$ and $Q_2=1-P_2$. $\phi_{q}(0)$ is the zero time value
of $\phi_{q}(t)$ and can be approximated by~\cite{Gotze76a}
$\frac{2S_{q}} {\beta \hbar \Delta n(\Omega_{q}) \Omega_{q}}$ where
$S_{q}$ is the static structure factor, $\Delta n(\omega) =
n(\omega)-n(-\omega)$ and $n(\omega)=\frac{1}{e^{\beta\hbar\omega}-1}$
is the Bose distribution function at temperature $T$.
\subsection{Quantum Mode-Coupling Approach}
We employ a quantum mode-coupling approach recently described by us
for quantum liquids~\cite{Kletenik11} to obtain the memory kernel
described by Eq.~(\ref{eq:memory}). This approach is based on our
early work to describe density fluctuations and transport in quantum
liquids such as liquid {\em para}-hydrogen, {\em ortho}-deuterium, and
normal liquid
helium.\cite{Reichman01a,Rabani02a,Reichman02a,Rabani02b,Rabani02c,Rabani02d,Rabani04,Rabani05,Rabani05a}
The basic idea behind this approach is that the random force projected
correlation function, which determines the memory kernel for the
intermediate scattering function, decays at intermediate and long
times predominantly into modes which are associated with
quasi-conserved dynamical variables. It is reasonable to assume that
the decay of the memory kernel at long times will be governed by those
modes that have the longest relaxation time. Thus, the first
approximation made by the QMCT is to
replace the projected time evolution operator, $\mbox{e}^{i
{\bar{\mathcal L}} t}$, by its projection onto the subspace spanned
by these slow modes.\cite{Rabanireview05} The second approximation
involves the factorization of four-point density correlations into a
product of two-point density correlation.\cite{Rabanireview05}
Following the derivation outlined by G{\"{o}}tze and L\"{u}cke (GL)
for zero temperature,\cite{Gotze76a,Gotze76b} the memory kernel at
finite temperature (in frequency space),$\tilde{M}_{q}(\omega) =
\int_{-\infty}^{\infty} dt \mbox{e}^{-i \omega t} M(q,t)$), can be
approximated by
\begin{eqnarray}
\tilde{M}_{q}(\omega) &\approx& \frac{\hbar m \beta^{2}}{4\pi\omega
q^{2} n} \int \frac{d^{3}k}{(2\pi)^3} V_{q,k,q-k}^{2}
\int_{-\infty}^{\infty}d\omega' \omega' \\ \nonumber & & \times
(\omega-\omega') T(\omega',\omega-\omega') \tilde{\phi}_{q-k}(\omega')
\tilde{\phi}_{k} (\omega-\omega'),
\label{eq:Mqmct}
\end{eqnarray}
where $n$ is the number density,
$\tilde{\phi}_{q}(\omega)=\int_{-\infty}^{\infty} dt \mbox{e}^{i
\omega t} \phi_{q}(t)$ is the Fourier transform of the Kubo
transform of the intermediate scattering function and
\begin{eqnarray}
T(\omega_{1},\omega_{2})
= n(-\omega_{1}) n(-\omega_{2}) - n(\omega_{1}) n(\omega_{2}).
\label{eq:T}
\end{eqnarray}
The vertex, $V_{q,k,q-k}$, is formally given by
\begin{eqnarray}
N_{q-k,k} V_{q,k,q-k} &=& \left(Q{\mathcal{L}}^{2}\hat{\rho}_{q}
|\hat{\rho}_{k}\hat{\rho}_{q-k}\right) \\ \nonumber &=&
\left(\mathcal{L}^{2}\hat{\rho}_{q}|\hat{\rho}_{k}\hat{\rho}_{q-k}\right)
- \Omega_{q}^{2} \left(\hat{\rho}_{q}| \hat{\rho}_{k}
\hat{\rho}_{q-k}\right),
\label{eq:vertex}
\end{eqnarray}
with the normalization approximated by
\begin{eqnarray}
N_{q-k,k} &=& \left( \hat{\rho}_{q} \hat{\rho}_{q-k}| \hat{\rho}_{q-k}
\hat{\rho}_{k} \right)(0) \\ \nonumber &\approx& \hbar \beta
\int_{-\infty}^{\infty} \frac{d\omega} {\pi} \int_{-\infty}^{\infty}
\frac{d\omega'} {\pi} \frac{1} {4\omega}T(\omega',\omega-\omega')
\\ \nonumber & & \times \omega'(\omega-\omega')
\tilde{\phi}_{q-k}(\omega') \tilde{\phi}_{k}(\omega-\omega'),
\label{eq:normalization}
\end{eqnarray}
consistent with the spirit of QMCT where four-point density
correlations are factorized into a product of two-point density
correlations.\cite{Rabanireview05}
\subsection{The vertex}
The vertex in Eq.~(\ref{eq:vertex}) is difficult to compute since it
involves three-point Kubo density correlations. A common approach
taken by classical mode-coupling theory (CMCT) is based on a
convolution approximation.\cite{Jackson62} For the Kubo transform
quantum case, a convolution-like approach is not unique. The approach
we adopt here is based on an extension of the work of GL to finite
temperatures.\cite{Gotze76a,Gotze76b} In this work, a dynamical
approximation is made to remove the dependence on Kubo transformed
structure factor in the vertex. The assumption behind this
approximation is that the major contribution to the vertex and its
normalization comes from a characteristic frequency of the
system. Thus, we approximate $\tilde{\phi}_{q}(\omega)$ within the
vertex by
\begin{equation}
\tilde{\phi}_{q}(\omega)=\frac{2\pi S_{q}}{\beta \hbar \Delta
n(\Omega_{q}) \omega}
(\delta(\omega-\Omega_{q})-\delta(\omega+\Omega_{q})).
\label{eq:phiqw-approx}
\end{equation}
which satisfies the known sum rule $\int_{-\infty}^{\infty} d\omega
\tilde{\phi}_{q}(\omega) = \phi_{q}(0)$. Inserting this approximation
for $\tilde{\phi}_{q}(\omega)$ into the expression for $N_{q-k,k}$
given by Eq.~(\ref{eq:normalization}) yields:
\begin{equation}
N_{q-k,k} \approx \frac{2S_{q-k}S_{k}} {\hbar\beta\Delta
n(\Omega_{q-k})\Delta n(\Omega_{k})} K(\Omega_{q-k},\Omega_{k}),
\label{eq:N-approx}
\end{equation}
where
\begin{equation}
K(\Omega_{q-k},\Omega_{k}) = \frac{T(\Omega_{q-k},\Omega_{k})}
{\Omega_{q-k}+\Omega_{k}} + \frac{T(-\Omega_{q-k},\Omega_{k})}
{\Omega_{q-k}-\Omega_{k}}.
\label{eq:K}
\end{equation}
For $V_{q,k,q-k}$ we use the exact relations~\cite{Kletenik11}
\begin{equation}
\left(L^{2}\hat{\rho}_{q} |\hat{\rho}_{k} \hat{\rho}_{q-k}\right) =
\frac{1}{m\beta}\left(q\cdot k S_{q-k}+q\cdot(q-k)S_{k}\right).
\label{eq:gl1}
\end{equation}
and the convolution approximation
\begin{equation}
\left<\hat{\rho}_{q}^{\dagger}(t),\hat{\rho}_{k}\hat{\rho}_{q-k}\right>\approx
S_{q}S_{k}S_{q-k}
\label{eq:convGL}
\end{equation}
to obtain the approximation to the vertex:
\begin{widetext}
\begin{eqnarray}
V_{q,k,q-k} &=& \frac{\Delta n(\Omega_{q-k})\Delta n(\Omega_{k})
C_{q,k,q-k}}{S_{q-k}S_{k}K(\Omega_{q-k},\Omega_{k})}
\left[\frac{(\Omega_{k} +\Omega_{q-k})^{2}-\Omega_{q}^{2}}
{(\Omega_{k}+\Omega_{q-k})}\right]
\label{eq:vertexGL}
\end{eqnarray}
where
\begin{eqnarray}
C_{q,k,q-k}&=&\frac{\Omega_{q}S_{q}S_{k}S_{q-k}-\frac{\hbar \Delta
n(\Omega_{q})}{2m}\left[q\cdot kS_{q-k}+q\cdot(q-k)S_{k}\right]}
{\Omega_{q}\Delta n(\Omega_{k}+\Omega_{q-k})
-(\Omega_{k}+\Omega_{q-k})\Delta n(\Omega_{q})}
\label{eq:C}
\end{eqnarray}
\end{widetext}
The above expressions close the equation of motion
(Eq.~(\ref{eq:qgle})) and require only the static structure factor to
produce a full approximation to the time dependence of the quantum
density-density time autocorrelation function.
\subsection{High and Low temperature limits}
It may be shown that the above equations reduce to the
venerable classical mode-coupling equations in the high temperature
limit and to the GL theory as $T \rightarrow 0$. The latter theory
produces a representation of the dispersion of superfluid helium that
is at least as accurate as the Feynman-Cohen (FC)
theory~\cite{Feynman56} at low values of $q$ and exhibits
Pitaevskii-bending of the spectrum at high $q$, unlike the FC theory.
In particular at high $T$,
\begin{eqnarray}
\lim_{\beta\rightarrow0}M_{q}(t)&\approx&\frac{k_{\rm B}T
n}{16\pi^{3}m q^2} \int d^{3}k \left({q}\cdot k c_{k}
\right.\\ \nonumber &+& \left. {q} \cdot(q-k) c_{q-k}\right)^{2}
\phi_{q-k}(t)\phi_{k}(t),
\label{eq:Mqtclassical}
\end{eqnarray}
where $c_{q} = \frac{1}{n} \left(1-\frac{1}{S_{q}} \right)$ is the
direct correlation function. In addition, $\phi_{q}(t)$ reduces to
the classical intermediate scattering function, $F(q,t)$ as $\beta \to
0$. This is recognized as the CMCT memory function.\cite{Gotze09}
At $T \rightarrow 0$ the equation for the memory function reduces to:
\begin{eqnarray}
\lim_{T\rightarrow0}\tilde{M}_{q}(\omega)&\approx&\frac{\hbar
m\beta^{2}}{2 n \omega q^{2}} \int \frac{d^{3}k}{(2\pi)^{3}}
V_{q,k,q-k}^{2} \\ \nonumber & & \times \int_{0}^{\omega}
\frac{d\omega'}{\pi}\omega'(\omega-\omega')
\tilde{\phi}_{q-k}(\omega')\tilde{\phi}_{k}(\omega-\omega'),
\label{eq:MqtT=0}
\end{eqnarray}
with
\begin{eqnarray}
\lim_{T\rightarrow0}V_{q,k,q-k} = & &\frac{\hbar n}{2m} (\omega_{k} +
\omega_{q-k} + \omega_{q}) \\ && \nonumber \times \left(q\cdot k c_{k}
+ q\cdot(q-k)c_{q-k}\right),
\label{eq:vertexT=0}
\end{eqnarray}
which are the $T \rightarrow 0$ equations for quantum density
fluctuations in superfluid helium first derived by
GL.\cite{Gotze76a,Gotze76b} In the above, $\omega_q=\frac{\hbar
q^2}{2mS_q}$. We note in passing that the term $\beta^2$ appearing in
Eq.~(\ref{eq:MqtT=0}) (and not in the derivation of GL) arises from
our definition of the Kubo transform (Eq.~\ref{eq:phiqt-def}), which
includes a $\frac{1}{\hbar \beta}$, while that of GL does not. Care
must be taken applying the Kubo transform as $T \rightarrow 0$.
In the $T \rightarrow 0$ case, the entire structure of the memory
function differs greatly from that of its high temperature counterpart
and the convolution structure is lost. Eqs.(\ref{eq:MqtT=0}) and
(\ref{eq:vertexT=0}) do not imply a memory function that is a product
of correlators at identical times. This is a consequence of the
quantum fluctuation-dissipation theorem (QFDT) that must be
satisfied. At $T \rightarrow 0$ the function $T(\omega_q,\omega_k)$
becomes proportional to the difference of a product of step-functions
in frequency, dramatically altering the structure of the theory. This
distinction between the low and high temperature limits has important
consequences, as discussed below.
\subsection{Nonergodic parameter}
The nonergodic parameter,
\begin{equation}
f_{q}=\frac{\phi_{q}(t\rightarrow\infty)}{\phi_{q}(0)}=\frac{\hbar
\Delta
n(\Omega_{q})\Omega_{q}}{2k_{B}TS_{q}}\phi_{q}(t\rightarrow\infty),
\label{eq:f}
\end{equation}
is often used to describe the ergodic to nonergodic transition as the
liquid is cooled down to the mode-coupling critical temperature
$T_c$. Above $T_c$ one finds a single solution where $f_{q}=0$ for all
values of $q$, while at $T_c$ the nonergodic parameter acquires a
finite value $f_{q} > 0$.\cite{Gotze92} It is simple to show that
$f_{q}$ must satisfy the equation:\cite{BalucaniZoppi}
\begin{equation}
\frac{f_{q}}{1-f_{q}}=\frac{1}{\Omega_{q}^{2}} M_{q}(t \rightarrow
\infty)
\label{eq:nep}
\end{equation}
The above equation for the nonergodic parameter reflects the structure
of the QGLE (Eq.~(\ref{eq:qgle})), and thus, is valid both in the
classical and quantal limits. In the former, the long time limit of
the memory kernel is given by $M_{q}(t\rightarrow\infty) \approx
\frac{k_{\rm B}T n}{16\pi^{3}m q^2} \int d^{3}k {\bar V}_{q,k,q-k}^{2}
f_{q-k} f_{k}$ with ${\bar V}_{q,k,q-k}^{2} = S_{q-k}S_{k}
\left({q}\cdot k c_{k} + {q} \cdot(q-k) c_{q-k}\right)^{2}$. The
quantum case is a bit more complicated since the structure of the
memory kernel is quite different and involves a convolution of
products of $\tilde{\phi}_{q}(\omega)$. The derivation for
$M_{q}(t\rightarrow\infty)$ is thus, based on the following expansion:
\begin{widetext}
\begin{equation}
\begin{split}
\frac{1}{\omega}T(\omega',\omega-\omega') & \omega'(\omega-\omega') =
\frac{1}{\beta \hbar} + \frac{\beta \hbar}{12}\omega'(\omega-\omega') -
\frac{(\beta \hbar)^{3}}{720}\left(\omega'^{2}-\omega'(\omega-\omega') +
(\omega-\omega')^{2}\right)+\\ & \frac{(\beta \hbar)^{5}}{30240}
\left((\omega-\omega')^{4} - \omega'(\omega-\omega')^{3} +
\omega'^{2}(\omega-\omega')^{2} -
\omega'^{3}(\omega-\omega')+\omega'^{4}\right)+O(\beta^{7}).
\end{split}
\label{eq:expand}
\end{equation}
\end{widetext}
Inserting this into the memory kernel (Eq.~(\ref{eq:Mqmct})) and
keeping the first two terms only, we obtain:
\begin{eqnarray}
\tilde{M}_{q}(\omega) &\approx& \frac{\hbar m \beta^{2}}{4\pi q^{2} n}
\int \frac{d^{3}k}{(2\pi)^3} V_{q,k,q-k}^{2}
\int_{-\infty}^{\infty}d\omega' \left(\frac{1}{\beta \hbar} + \right. \\ & &
\nonumber \left. \frac{\beta \hbar}{12}\omega' (\omega-\omega') +
\cdots\right) \tilde{\phi}_{q-k}(\omega') \tilde{\phi}_{k}
(\omega-\omega').
\label{eq:Mqmct1}
\end{eqnarray}
In the time domain, this translates to:
\begin{eqnarray}
M_{q}(t) &\approx& \frac{\hbar m \beta^{2}}{2 q^{2} n} \int
\frac{d^{3}k}{(2\pi)^3} V_{q,k,q-k}^{2} \\ &&
\nonumber \times \left(\frac{1}{\beta \hbar} \phi_{q-k}(t) \phi_{k}(t) +
\frac{\beta \hbar}{12}\dot{\phi}_{q-k}(t)\dot{\phi}_{k}(t) + \cdots\right),
\label{eq:Mqmct2}
\end{eqnarray}
where the dot denotes a time derivative, i.e., $\dot{\phi}_{k}(t) =
\frac{d\phi_{k}(t)}{dt}$. The other terms in the expansion of
Eq.~(\ref{eq:expand}) that have been omitted give rise to terms of the
form
\begin{equation}
\sum_{j}a_{j}(\beta)\phi_{q-k}^{(j)}(t)\phi_{k}^{(n-j)}(t),
\end{equation}
where $a_{j}(\beta)$ are related to the expansion coefficients of
$\frac{1}{\omega}T(\omega',\omega-\omega')\omega'(\omega-\omega')$
and $\phi_{k}^{(j)}(t)=\frac{d^{j}\phi_{k}(t)}{dt^{j}}$ is the $j$'s
time derivative of $\phi_{k}(t)$.
The long time limit of the Eq.~(\ref{eq:Mqmct2}) is now given by:
\begin{eqnarray}
M_{q}(t \rightarrow \infty) &\approx& \frac{m \beta}{2 q^{2}
n} \int \frac{d^{3}k}{(2\pi)^3} V_{q,k,q-k}^{2} \\ & & \nonumber
\times \phi_{q-k}(t \rightarrow \infty) \phi_{k}(t \rightarrow
\infty),
\label{eq:Mqmct3}
\end{eqnarray}
where all the time derivatives vanish as $t \rightarrow \infty$ even
when $\phi_{k}(t \rightarrow \infty)$ decays to a constant. Finally,
we can rewrite the above in terms of the nonergodic parameter:
\begin{eqnarray}
M_{q}(t \rightarrow \infty) &\approx& \frac{m \beta}{2 q^{2}
n} \int \frac{d^{3}k}{(2\pi)^3} \\ & & \nonumber \times
V_{q,k,q-k}^{2} \phi_{q-k}(0) \phi_{k}(0) f_{q-k} f_{k}.
\label{eq:Mqmct4}
\end{eqnarray}
The above expression is strictly valid at $T \rightarrow 0$ but not at
$T=0$, since the expansion given by Eq.~(\ref{eq:expand}) is not valid
at $T=0$. The final result is similar to the classical equation,
however the vertex is given by the full quantum mechanical expression
of Eq.~(\ref{eq:vertexGL}).
\section{Quantum integral equation theory}
\label{sec:qiet}
The QMCT requires as input the static structure factor, $S_q$ and its
Kubo transform $\phi_q(0)$. Here, instead of using PIMC to generate
this input,\cite{Rabani01a} we refer to a quantum integral equation
approach, that is based on the early work of Chandler and
Richardson.\cite{Chandler84a,Chandler84b} We begin with the
Ornstein-Zernike relation applicable to quantum liquids. The quantum
system composed of $N$ particles can be mapped on a classical system
consisting of $N$ ring polymers, each polymer being composed of $P$
beads. Then, we can write the matrix RISM (reference interaction site
model~\cite{Chandler84a,Chandler84b}) equation for the classical
isomorphic system by:
\begin{equation}
h(|{\bf r}-{\bf r'}|) = w * c * w(|{\bf r}-{\bf r'}|) + n
w * c * h(|{\bf r}-{\bf r'}|),
\label{eq:QOZ}
\end{equation}
where $*$ denotes a convolution integral and as before, $n$ is the
number density. In the above equation, $h(r)$, $w(r)$, and
$c(r)$ are the total correlation function, the self correlation
function, and direct correlation function, respectively, defined by:
\begin{equation}
\begin{split}
h(r) = \frac{1}{\hbar\beta}\int_0^{\hbar\beta} d\lambda h(r,\lambda) \\
w(r) = \frac{1}{\hbar\beta}\int_0^{\hbar\beta} d\lambda w(r,\lambda) \\
c(r) = \frac{1}{\hbar\beta}\int_0^{\hbar\beta} d\lambda c(r,\lambda), \\
\end{split}
\label{eq:hwc}
\end{equation}
and $h(r,\lambda)$, $w(r,\lambda)$, and $c(r,\lambda)$ are the
imaginary time total, self, and direct correlation functions,
respectively. In the classical limit Eq.~(\ref{eq:QOZ}) reduces to the
classical Ornstein-Zernike equation with $w(r)=1$. In what
follows, we will use the notation $\tilde{w}_q(\lambda)$ for the
Fourier transform of $w(r,\lambda)$, and similarly for
$\tilde{c}_q(\lambda)$ and $\tilde{h}_q(\lambda)$:
\begin{equation}
\begin{split}
\tilde{h_q} = \frac{1}{\hbar\beta}\int_0^{\hbar\beta} d\lambda
\tilde{h}_q(\lambda) \\ \tilde{w}_q =
\frac{1}{\hbar\beta}\int_0^{\hbar\beta} d\lambda
\tilde{w}_q(\lambda) \\ \tilde{c}_q =
\frac{1}{\hbar\beta}\int_0^{\hbar\beta} d\lambda \tilde{c}_q(\lambda).
\\
\end{split}
\label{eq:hwcq}
\end{equation}
To proceed, we refer to the mean-pair interaction approximations along
with the quadratic reference action~\cite{Chandler84a} and rewrite:
\begin{equation}
\tilde{w}_q(\lambda) = \exp\{ -q^2 R^2(\lambda)\},
\label{eq:wqlambda}
\end{equation}
where
\begin{equation}
R^2(\lambda) = \sum_j \frac{1-\cos(\Omega_j \lambda)}{\beta m
\Omega_j^2 + \alpha_j},
\label{eq:R}
\end{equation}
$m$ is the particle mass, $\Omega_j=2\pi j/\hbar \beta$ is the
Matsubara frequency and $\alpha_j$ is given by:
\begin{equation}
\alpha_j = \frac{1}{6 \pi^2 \hbar \beta} \int_0^{\infty} dq
\int_0^{\hbar \beta} d\lambda q^4 \tilde{v_q} (1-\cos(\Omega_j
\lambda) \tilde{w}_q(\lambda).
\label{eq:alpha}
\end{equation}
In the above the solvent induced self-interaction is given by:
\begin{equation}
\tilde{v_q} = -\tilde{c}_q^2(n \tilde{w}_q + n^2 \tilde{h}_q).
\label{eq:v}
\end{equation}
In order to close the quantum Ornstein-Zernike equations, which in
$q$-space can be written as:
\begin{equation}
\tilde{h}_q = \tilde{w}_q \tilde{c}_q \tilde{w}_q + n
\tilde{w}_q \tilde{c}_q \tilde{h}_q,
\label{eq:QOZq}
\end{equation}
we use the Percus-Yevick (PY) closure of the form (in $r$-space):
\begin{equation}
c(r)=(h(r)+c(r)+1) (\exp(-\beta v(r)) - 1),
\label{eq:PY}
\end{equation}
where $v(r)$ is the pair interaction between two particles. The static
structure factor and its Kubo transform are then given by:
\begin{eqnarray}
S_q = 1 + n \tilde{h}_q\\ \nonumber
\phi_q(0) = \tilde{w}_q + n \tilde{h}_q.
\label{eq:sqfromhq}
\end{eqnarray}
In all the applications reported below we have used the approximate
relation for $\phi_{q}(0) \approx \frac{2S_{q}} {\beta \hbar \Delta
n(\Omega_{q}) \Omega_{q}}$.
\section{Ring polymer molecular dynamics}
\label{sec:rpmd}
The RPMD approach to quantum dynamics provides an approximation to
quantum mechanical Kubo transformed correlation functions by using a
classical evolution of the imaginary time
paths~\cite{Manolopoulos04}. Consider a multidimensional system of $N$
distinguishable particles with a Hamiltonian of the form,
\begin{equation}
H = \sum_{\alpha=1}^{N} \frac{{\bf p}_{\alpha}^{2}}{2m_\alpha}+V({\bf
r}_1,\ldots,{\bf r}_N),
\label{eq:rpmd1}
\end{equation}
where, ${\bf r}_{\alpha}$ and ${\bf p}_{\alpha}$ are the positions and
momenta of the particles and $V({\bf r}_1,\ldots,{\bf r}_N)$ is the
potential energy of the system. The RPMD approximation to the
canonical correlation function, $\tilde{c}_{AB}(t)$, for position
dependent operators $A({\bf r})$ and $B({\bf r})$ is,
\begin{eqnarray}
\tilde{c}_{AB}(t) & \simeq & \frac{1}{(2\pi\hbar)^{3NP}Z_P}\int
d^{3NP}{\bf p}\int d^{3NP}{\bf r} \nonumber \\ & & \,e^{- \beta_{P}
H_P({\bf p},{\bf r})}A_P({\bf r})B_P({\bf r}_t),
\label{eq:cab}
\end{eqnarray}
where
\begin{equation}
Z_{P} = \frac{1}{(2\pi\hbar)^{3NP}}\int d^{3NP}{\bf p}\int d^{3NP}{\bf
r}~ \,e^{-\beta_{P} H_{P}({\bf p},{\bf r})},
\end{equation}
and $\beta_{P} =\beta/P$. $H_P({\bf p},{\bf r})$ is the classical
Hamiltonian of the $N$ particle $P$ bead ring polymers with the
external potential of $V({\bf r}_1,\ldots, {\bf r}_N)$ acting on each
bead,
\begin{eqnarray}
H_P({\bf p},{\bf r}) & = & \sum_{\alpha=1}^{N} \sum_{k=1}^{P}
\left(\frac{({\bf
p}_{\alpha}^{(k)})^2}{2m_i}+\frac{1}{2}m_{\alpha}\omega_P^2 ({\bf
r}_{\alpha}^{(k)}-{\bf r}_{\alpha}^{(k+1)})^2\right) \nonumber \\ &&
+\sum_{k=1}^{P} V({\bf r}_1^{(k)},\ldots,{\bf r}_N^{(k)}),
\label{eq:hp}
\end{eqnarray}
where $\omega_{P}=1/\beta \hbar$ and the cyclic boundary condition
${\bf r}_{\alpha}^{(P+1)} \equiv {\bf r}_{\alpha}^{(1)}$ applies. The
time-evolved coordinates ${\bf r}_t\equiv {\bf r}_t({\bf p},{\bf r})$
in Eq.~(\ref{eq:cab}) are obtained from the classical dynamics
generated from this Hamiltonian and the operators $A_P({\bf r})$ and
$B_P({\bf r}_t)$ are evaluated by averaging over the beads of the ring
polymer at times $0$ and $t$ respectively,
\begin{equation}
A_P({\bf r}) = \frac{1}{P}\sum_{k=1}^P A({\bf r}_1^{(k)},\ldots,{\bf r}_N^{(k)}),
\label{eq:rpmd11}
\end{equation}
\begin{equation}
B_P({\bf r}) = \frac{1}{P}\sum_{k=1}^P B({\bf r}_1^{(k)},\ldots,{\bf r}_N^{(k)}).
\label{eq:rpmd12}
\end{equation}
The RPMD method has previously been used to study a diverse selection
of multidimensional systems including proton transfer between organic
molecules,\cite{Collepardo-Guevara2008} diffusion in and inelastic
neutron scattering from liquid
para-hydrogen,\cite{Miller2005,Craig2006} diffusion of light atoms in
liquid water,\cite{Markland2008} and gas phase reactions such as that
between methane and hydrogen.\cite{Suleimanov2011} In all cases RPMD
has been able to capture the dominant quantum mechanical effects in
the dynamics and provide good agreement with the available
experimental or exact results. RPMD has also been applied to look at
deep tunneling of Muonium and Hydrogen atoms in
ice~\cite{Markland2008} and in this regime has been shown to be
related to semi-classical Instanton theory.\cite{Richardson09}
\section{Simulations Details}
\label{sec:sim}
\begin{figure*}[t]
\includegraphics[width=16cm]{phase}
\caption{Panel~(a.): The diffusion constant of particles of type
A as a function of the quantumness, $\Lambda^{*}$, obtained from the
RPMD simulations for a quantum Kob-Anderson LJ binary mixture for
two temperatures. Panel~(b.): Dynamic phase diagram (volume
fraction versus quantumness) calculated from the QMCT for a
hard-sphere fluid. Panel~(c.): The mean square displacement of A
particles as obtained from the RPMD simulations for the classical
case (left frame, $\Lambda^{*}=0$), the trapped regime (middle
frame, $\Lambda^{*}=1.125$), and the regime governed by strong
quantum fluctuations (right frame, $\Lambda^{*}=1.1325$).}
\label{fig:phase}
\end{figure*}
The quantum mode coupling theory requires as input the static
structure factor and its Kubo transform. In the present study, we used
a single component hard sphere (HS) model to generate this input
within the frame work of the integral equation approach described
above. Using the PY closure, the system remains disordered even at
very high volume fraction, thus providing a simple model to explore
the quantum glass transition. The integral equations
(\ref{eq:hwcq})-(\ref{eq:alpha}) were solved self-consistently. A
simple trapezoidal integration scheme over the imaginary time axis was
employed, with $P=400$ slices (we have checked convergence of the
static input with respect to $P$). Here, $P$ is analogous to the
number of beads in the RPMD approach. For the HS system, it can be
shown that the quantum mode coupling equations scale with the ratio of
the de Broglie thermal wavelength to the particle size,
$\Lambda^{*}=\sqrt{\beta \hbar^2/m \sigma^2}$. Thus, to change the
quantumness, one can either change $\hbar$, or the mass, or the
temperature. For the QMCT results shown below, we have varied the
temperature to reflect a change in $\Lambda^{*}$. We note that the
temperature has no effect on the static structure factor in the
classical case.
\begin{table}
\begin{tabular*}{0.5\textwidth}{@{\extracolsep{\fill}}ccc}
\hline\hline
\multicolumn{1}{c}{Parameter} & \multicolumn{1}{c}{LJ units} & \multicolumn{1}{c}{Atomic Units} \\
\hline
$\epsilon_{AA}$ & 1 & 3.8x10$^{-4}$\\
$\epsilon_{BB}$ & 0.5 & 1.9x10$^{-4}$\\
$\epsilon_{AB}$ & 1.5 & 5.7x10$^{-4}$\\
$\sigma_{AA}$ & 1 & 6.43 \\
$\sigma_{BB}$ & 0.88 & 5.65 \\
$\sigma_{AB}$ & 0.8 & 5.14 \\
Mass$_{A}$ & 1 & 3646 \\
Mass$_{B}$ & 1 & 3646 \\\hline \hline
\end{tabular*}
\caption[]{Parameters used in our RPMD simulations on the Andersen-Kob
Lennard-Jones glass forming system.}
\label{ta:param}
\end{table}
We performed RPMD simulations on the Kob-Andersen glass forming
system,~\cite{Kob95a,Kob95b} a binary LJ fluid, because the HS system
investigated above by means of QMCT crystallizes on the timescale of
the RPMD simulations. Each simulation consisted of $1000$ particles,
$800$ of type A and $200$ of type B in a cubic box of length
$9.4\sigma_{AA}$. The LJ parameters are given in
Table~\ref{ta:param}.The equations of motion were integrated using a
time step of 0.005 in Lennard-Jones (LJ) units using the normal mode
integration scheme of Ref.~\onlinecite{Ceriotti2010}. The number of
beads, $P$, used was given by the formula,
\begin{equation}
P = \frac{11.2 \hbar}{T^*}.
\end{equation}
This choice gives good convergence for all the regimes studied.
Initial configurations were generated by annealing from a temperature
$T^{*}=5.0$ to the target temperature over a period of $2\times 10^6$
time-steps. From these initial configurations we ran a further
$2\times 10^5$ steps of equilibration using a targeted Langevin
equation normal mode thermostatting scheme.\cite{Ceriotti2010} This
was followed by microcanonical dynamics for $2\times 10^6$ steps
during which the results were collected. The quantum effect,
$\Lambda^{*}$, was varied by changing the parameter $\hbar$. Five
simulations were run for each temperature and value of $\hbar$ and the
results averaged.
\section{Results}
\label{sec:res}
Fig.~\ref{fig:phase} shows the results obtained from our QMCT
treatment of hard spheres and RPMD simulations of the KA binary LJ
fluid as the size of quantum fluctuations in the system are
varied.\cite{Markland11} Both of these systems have previously been
shown to exhibit all of the features of glassy behavior present in
more complex fluids. In panel~(b.) we show the liquid-glass dynamic
phase diagram that is obtained from the QMCT calculation. The phase
boundary is defined as the point where the solution of equations
Eqs.~(\ref{eq:f}), (\ref{eq:nep}) and (\ref{eq:Mqmct4}) leads to a
finite value for the nonergodic parameters, $f_q$. At this point QMCT
predicts that the system will never fully relax on any time-scale at
the given packing fraction. For the RPMD calculations, which are based
on the evolution of semi-classical trajectories, we instead show the
effect of quantum fluctuations on the diffusion coefficient of the
particles at two different temperatures ($T^{*}=2.0$ and $0.7$) as the
classical glass transition temperature of the system is approached
($T^{*} \approx 0.45$) in panel~(a.) of Fig. \ref{fig:phase}. Since
the mean square displacement of the particles in the ring polymer
trajectories show a caging regime (see the panel~(c.) of
Fig.~\ref{fig:phase}), the diffusion constant was extracted from the
long time slope of the mean-square displacement where the diffusive
regime had been reached. The size of the quantum fluctuations were
controlled by varying $\Lambda^{*}$, the ratio of the de Broglie
thermal wavelength to the particle size which controls the scale of
quantum behavior.
Comparing the RPMD results in panel~(a.) and QMCT results in
panel~(b.) of Fig.~\ref{fig:phase}, a remarkably consistent picture
emerges from these two different approaches to quantum dynamics and
glass forming systems. In the classical limit ($\Lambda^{*}
\rightarrow 0$) RPMD reduces to classical mechanics and QMCT to
classical MCT. As small quantum fluctuations are initially introduced,
little difference is observed in either the RPMD diffusion coefficient
or QMCT liquid-glass line. However, as $\Lambda^{*}$ is increased
beyond 0.1, quantum effects are at first found to promote and then
inhibit glass formation. In the case of RPMD, this is characterized by
a decrease of nearly three orders of magnitude in the diffusion
coefficient, and for QMCT, a 20 \% fall in the packing fraction
required for vitrification. When the thermal wavelength is increased
further and becomes on the order of the particle size, the diffusion
coefficient in the quantum system exceeds that observed in the
classical limit. In addition the RPMD simulations at $T^{*}=0.7$ and
$2.0$ indicate that size of the re-entrance becomes much larger as the
glass transition temperature is approached. Moreover, there is a hint
of an interesting effect where, at high values of $\Lambda^{*}$, the
diffusion coefficient at lower temperature exceeds that at the higher
temperature. We will return to this point later.
\begin{figure}[t]
\includegraphics[width=8cm]{gr}
\caption{The bead (upper panel) and centroid (lower panel) radial
distribution functions of A particles for a classical
($\Lambda^{*}=0$, dashed) and trapped quantum ($\Lambda^{*}=0.75$,
solid) regime. The bead distribution suggests less order in the
trapped regime compared to a classical simulation while the centroid
structure shows an increase in order.}
\label{fig:gr}
\end{figure}
Since both MCT and our new QMCT approach use the structure factor as
input it is instructive to see if the dynamical reentrance is hinted
at in this property. Fig.~\ref{fig:gr} shows the radial distribution
function (RDF), which is the spatial Fourier transform of the
structure factor, that has been calculated from the RPMD simulations
of the KALJ fluid. For static equilibrium properties such as the RDF,
RPMD gives numerically exact results since it reduces to the path
integral molecular dynamics approach.\cite{Parrinello1984} The true
(observable) quantum RDF is determined by the ring polymer bead
correlations and is shown in the top panel for both the classical
limit ($\Lambda^{*}=0$) and for a trapped regime
($\Lambda^{*}=0.75$). As quantum effects are introduced the RDF
exhibits a broadening of the peaks due to the increasing uncertainty
in the particle positions which acts to smear out the pair
structure. Throughout the entire range of $\Lambda^{*}$ studied the
structure is observed simply to broaden systematically with
$\Lambda^{*}$ and thus, there is no indication of the observed
dynamical reentrance in the RDF.
In the bottom panel we show the centroid RDF in which the centers of
the imaginary time paths, rather than the bead positions, were used to
compute the RDF. In the classical limit all beads collapse to a single
point and hence both ways of calculating the RDF become
identical. However as quantum fluctuations are increased the beads
spread further from the center of the polymer and hence the centroid
structure offers a different view into the structure of the quantum
liquid. Upon examining the centroid RDF in Fig ~\ref{fig:gr} one sees
the opposite trend upon increasing quantum fluctuations to that
observed in the bead RDF, i.e. weak quantum fluctuations lead to a
more structured centroid RDF which one would associate with more
glassy dynamics. As quantum fluctuations further increase this trend
reverses (data not shown). Hence the centroid pair distribution
function, which is not an experimental observable, appears to grossly
mimic the dynamical correlations observed in both the QMCT and RPMD
calculations. This is not entirely surprising, because one expects
that the centroid molecular dynamics (CMD) method,\cite{Voth96r} an
approach similar to RPMD, will also capture the reentrance. Since CMD
is an effective classical dynamics on the many-body centroid potential
and since there are situations where the many-body centroid potential
can be approximated by a sum of pair-wise potentials given by
$-k_{\tiny B}T \log g(r)$,\cite{Voth96,Voth04} such static
correlations in the centroid RDF must be evident if CMD is to
reproduce the same phenomenology as predicted by QMCT and RPMD. This
fact suggest that a strictly classical MCT calculation that uses a
static structure factor constructed from the centroid correlations
might be a good proxy for the full QMCT calculation. It should be
noted that the full QMCT only uses the observable structure factor and
thus one role played by the quantum vertex function is to effectively
convert the bead correlations to centroid ones via the quantum
fluctuation-dissipation theorem. The fact that the quantum vertex
involves frequency convolutions while the classical version does not
suggests, however, that there must be some distinction between a
classical MCT calculation with centroid correlations and the full
QMCT.
\begin{figure}[t]
\includegraphics[width=8cm]{rg}
\caption{Root-mean-square of the radius of gyration of A particles as
a function of $\Lambda^{*}$ obtained from the RPMD simulations for a
quantum Kob-Anderson LJ binary mixture for two temperatures. The
radius of gyration is defined as the average distance of the
replicas from the polymer center. The results are plotted for
temperatures $T^*=0.7$ (circles with dashed lines) and $T^*=2.0$
(triangles with dotted lines).}
\label{fig:rg}
\end{figure}
So what is the origin of the reentrance? For this we turn to the RPMD
trajectories to provide a physically insightful picture. Since this
approximation maps the dynamics of a quantum mechanical particles onto
that of a system of classical ring polymers, we can initially
interpret the results in the language of the diffusion of classical
polymers. In doing so we are careful to note that each bead on a given
polymer only interacts with the bead on another ring polymer
corresponding to the same imaginary time slice, a point which we will
return to later in this section. In the non-interacting limit, the
free ring polymer radius of gyration is directly proportional to the
thermal deBroglie wavelength of the quantum particle. Hence,
increasing $\Lambda^{*}$ allows the ring polymer representing each
quantum particle to spread out. The average radius of gyration of each
quantum particle in the interacting KALJ system is a static property
which can be calculated exactly from RPMD simulations. In
Fig.~\ref{fig:rg} we plot the average radius of gyration of each ring
polymer relative to the value in the free limit. The dependence of
this ratio on $\Lambda^*$ mimics the dependence of the diffusion
coefficient on $\Lambda^*$ shown in Fig.~\ref{fig:phase}. The decrease
in this ratio when reentrance is observed suggests a correlation
between the localization of the quantum particle and the increase in
the glassiness of the system. As quantum fluctuations are increased
from $\Lambda^{*} < 0.1$, the effective diameter of the quantum
particles differ little from $\sigma$ so that they can still fit into
the thermally accessible space, their radius of gyration is still well
approximated by $R_g^\text{free}$, and little change in the dynamics
is observed. However, once $\Lambda^{*}$ exceeds $0.1$ there is not
enough free space for the free ring polymers to further expand and
crowding due to the surrounding solvent cage causes the radius of
gyration to decrease from its free particle value.
In the upper panel of Fig.~\ref{fig:traj}, we show typical
configurations of a RPMD trajectory in the regime where the particle
is localized in a cavity. The particle is confined by its surrounding
neighbors, thus giving rise to an increase in its quantum kinetic
energy. For diffusion to occur, particles must push past each other,
causing further localization and incurring an even greater increase in
their kinetic energies. This energy penalty to motion leads to slower
dynamics. As $\Lambda^{*}$ is further increased, a tipping point is
reached when the thermal wavelength becomes comparable to the particle
size, $\Lambda^{*} \approx 1$. At this point, the cost of localization
becomes so large that the induced quantum kinetic energy enables the
crossing of barriers between cavities, leading to a rise in the radius
of gyration and facilitating diffusion. This can be seen in the
representative snapshots of a RPMD trajectory shown in the lower
panels of Fig.~\ref{fig:traj} in which the particle is delocalized
across two cavities. Accordingly, the radius of gyration recovers with
a corresponding increase in diffusion coefficient and diminishing of
the caging regime. This can be likened to a ``lakes to oceans''
percolation transition, in which the caging regime reflects
frustration of the quantum particle in the classical potential, a
frustration which is reduced when the kinetic energy of confinement
essentially floods the barriers and allows the particle to traverse
the region between adjacent potential energy minima.
\begin{figure}[t]
\includegraphics[width=8cm]{traj}
\caption{A series of snapshots taken from simulations at
$\Lambda^{*}=1.125$ (upper panels) and $\Lambda^{*}=1.3125$ (lower
panels) with $T^{*}=0.7$. For clarity the full imaginary time path
(colored red) is only shown for one particle of type A with all
others represented by their centroids. The centroids for the other
particles of types A and B are colored green and blue, respectively.
The upper panels depict configurations which reside in the trapped
regime where the ring polymer is essentially localized in one cavity
cage whereas in the tunneling regime (lower panels) it is frequently
spread across two or more cavities in the liquid resulting in more
facile motion.}
\label{fig:traj}
\end{figure}
Reentrant effects in quantum systems have also been observed in the
diffusion of electrons in a sea of classical random
blockers~\cite{Leung1994} as well as in model
systems.\cite{Wolynes03,Zamponi11} In the former case the problem can
be exactly mapped onto the diffusion of a classical ring
polymer. However, in our case, while the expression ``ring polymer''
is used to describe the isomorphism arising from the imaginary time
path integral representation described in
Eqs.~(\ref{eq:rpmd1})-(\ref{eq:hp}), it is not simply that of a system
of true harmonic ring polymers. This is because each bead of a polymer
only interacts with its corresponding bead at the same imaginary time
on the polymer representing another particle and not with any other
beads on that particle. One might therefore expect that in systems
with strong interactions it might be advantageous for the polymers to
correlate their beads so as to minimize repulsion in exchange for a
loss in entropy. To investigate this, we define vectors ${\bf
R}^{k}_{\alpha} ={\bf r}^{(k)}_{\alpha}-{\bf r}_{\alpha}^{c}$, which
represent the position of the bead at imaginary time $k$ on ring
polymer $\alpha$ relative to the position of the centroid (${\bf
r}_{\alpha}^c=(1/P)\sum_{k=1}^P{\bf r}_{\alpha}^{(k)})$, and we
define the angle between vectors ${\bf R}^k_{\alpha}$ and ${\bf
R}^{(k)}_{\beta}$ as,
\begin{equation}
\cos\theta^{(k)}_{\alpha,\beta} = \frac{{\bf R}^{(k)}_{\alpha} \cdot {\bf R}^{(k)}_{\beta}}{\mid {\bf R}^{(k)}_{\alpha} \mid \mid {\bf R}^{(k)}_{\beta} \mid}.
\end{equation}
This function, $\cos\theta^{(k)}_{\alpha,\beta}$, will have a value
of $-1$ if the $k$-th beads on polymers $\alpha$ and $\beta$ are aligned
perfectly away from each other and $+1$ if the beads are aligned
towards each other. Since any correlation between the beads on two
different particles is likely to be more pronounced at short distances
where interactions are stronger we plot the correlation function
$C(r)$,
\begin{equation}
C(r) = \langle \frac{1}{N}\sum_{\alpha>\beta}\frac{1}{P}
\sum_{k=1}^P\cos\theta^{(k)}_{\alpha,\beta} \delta(r-|{\bf
r}_{\alpha,\beta}^c|)\rangle,
\label{eq:cr}
\end{equation}
as a function of the distance between the centroids of two ring
polymers.
\begin{figure}[t]
\includegraphics[width=8cm]{c}
\caption{The bead vector correlation (see Eq.~(\ref{eq:cr})) for a
trapped regime with $\Lambda^{*}=0.75$ (left panel) and regime where
quantum fluctuations are pronounced with $\Lambda^{*}=1.3125$ (right
panel). The solid lines represent the bead vector correlations
between A particles and the dashed ones those between B particles.
In both cases $T^{*}=0.7$. In the trapped regime the ring polymer
beads show a large positive correlation around $r=\sigma$ which
results in a large repulsion when the particles attempt to move past
each other. In the other regime the beads align such that the
correlation is largely negative which facilitates particle motion.}
\label{fig:c}
\end{figure}
The function $C(r)$ is shown in Fig.~\ref{fig:c} for $\Lambda^{*}=0.75
$, which corresponds to the trapped regime, and for $\Lambda= 1.3125$,
which corresponds to the strong quantum fluctuation regime. For $r\leq
\sigma$ $C(r)$ is negative in both cases. At these distances the
potential between particles is strongly repulsive and hence for
polymers to approach this close their beads for the same imaginary
time must avoid each other. However for $r\approx \sigma$, $C(r)$
corresponding to the lower value of $\Lambda^{*}$, the correlation
becomes positive. At this distance the pair potential is attractive
and hence the energy of the system is lowered if the polymer arranges
its beads such that they are aligned on the same side of the
respective ring-polymers. However at the higher value of
$\Lambda^{*}$, the entropic cost of such an ordering outweighs the
energetic benefit, and hence $C(r)$ is negative. This coincides with
the change between the dynamical regimes of quantum trapping and
strong fluctuations because, for diffusion to occur, particles must
move past each other. This regime corresponds to enhanced tunneling.
In the case of low $\Lambda^{*}$ the beads of the polymer in the first
coordination shell at $r=\sigma$ are largely aligned such that pushing
them together induces a larger repulsion than if no such correlation
existed. This increases the barrier to diffusion in this regime.
One natural question that arises from this interpretation of our
results is what occurs if the RPMD calculations are carried out at
constant pressure rather than constant volume. The analogous QMCT
calculations are constant volume calculations and indeed, as far as
the hard-sphere control variables of volume fraction and $\Lambda^{*}$
are concerned, this question is irrelevant. The pressure varies as
the volume fraction, which can then just be rescaled to yield results
identical to those presented in the panel~(b.) of
Fig.~\ref{fig:phase}. However from the standpoint of thermal
variation, e.g. the variation of diffusion at fixed temperature while
varying $\Lambda^{*}$ (see panel~(a.) of Fig.~\ref{fig:phase}), this
question needs to be addressed. A natural expectation is that at
constant pressure the reentrant effect will be mitigated or destroyed
as the system can now adjust its volume as a natural response to the
buildup of local pressure created by the ``swelling'' of the ring
polymer. However, some aspects of this effect have been observed, for
example, in analogous reentrant-like effects seen in
Ref.~\onlinecite{Markland2008}, where quantization of a single species
in a constant pressure classical bath produces a reduction of the
effective diffusion constant. More generally the values of
$\Lambda^{*}$ for which the slowing of the liquid is observed are
highly realizable in room temperature systems. The thermal wavelength
of hydrogen at 300 K is 1.0 $\AA$ and hence the region of quantum
slowing corresponds to diffusion in a medium with particles of radius
2 to 5 $\AA$. Such a slow down is evident in experimental measurements
of the diffusion of hydrogen in non-glassy media such as water and
palladium.~\cite{Markland2008,Wipf1997}
We have carried out RPMD simulations of the binary glass-forming
system at constant pressure, and indeed found at least a strong
mitigation of the reentrant effect. Currently our statistics are not
sufficient to make definitive statements about dynamical behavior in
these systems, and thus these results will be reported in a future
publication. Regardless, it is clear that constant volume (confined)
systems will exhibit a strong enhancement of the effects reported
here. Further it should be mentioned that a similar reentrance is seen
in lattice models of quantum glasses where the concept of swelling of
imaginary time paths cannot be invoked to explain reentrant
relaxation.~\cite{Wolynes03,Zamponi11}
On a final note, a subtle feature of the RPMD results of
Fig.~\ref{fig:phase} (panel~(a.)) should be mentioned. At very large
values of $\Lambda^{*}$ the isothermal diffusion curves appear to
cross. While the effect is quite small, this crossing would imply a
reentrance of a different sort, namely a ``melting by cooling''
mechanism. This type of reentrance, distinct from that discussed for
the bulk of this work, is similar to that discussed in
Ref.~\onlinecite{Wolynes03}. It should be noted, however, that the
$\Lambda^{*}$ values here are large enough that particle statistics
cannot be neglected in the simulation of a realistic quantum fluid,
and the inclusion of such features may obviate this effect.
\section{Concluding remarks}
\label{sec:con}
In this work we have presented a self-contained discussion of
predictions for quantum glasses made by QMCT and RPMD. The predictions
of these two distinct, albeit highly approximate, theories appear to
be in harmony with each other. Both predict a strong reentrance in
the relaxation of quantum supercooled liquids, namely that weak
quantum fluctuations actually serve to push the system closer towards
the glass transition. This seemingly paradoxical effect has also been
noted in lattice models of quantum glasses and in models of quantum
optimization. Indeed, one interesting aspect of our work is that it
suggests that typical quantum annealing protocols should generically
have regions of parameter space where they are in fact less efficient
than their classical counterparts.
Future work will be directed towards the inclusion of bosonic
statistics into the formulation of QMCT so that an investigation of
the putative superglass may be carried out in a microscopic manner.
In addition, it would be interesting to investigate more complex
liquids such as confined supercooled water to see if quantum effects
which may manifest at high temperatures lead to novel dynamical
relaxation patterns. These topics will be reserved for the future.
\section{Acknowledgments}
The authors acknowledge Francesco Zamponi for useful discussions. KM
acknowledges support from Kakenhi grant No. 21015001 and 2154016. BJB
acknowledges support from NSF grant No. CHE-0910943. DRR would like to
thank the NSF through grant No. CHE-0719089 for support. ER and DRR
thank the US-Israel Binational Science Foundation for support.
\bibliographystyle{aip}
|
\section{Introduction}
Let $(M,g)$ be a smooth complete Riemannian manifold of dimension
$n$. A second-order linear differential operator
$L: C^\infty(M)\rightarrow C^\infty(M)$ without
zeroth order term can be written as
\begin{equation}\label{eq111}
Lf=Tr(A\circ hess(f))+g(V,\nabla f),
\end{equation}
where $A\in \Gamma(\textrm{End}(TM))$ is self-adjoint with respect
to $g$, $hess(f)\in \Gamma(\textrm{End} (TM))$ is the Hessian of
$f$ in the form defined by $hess(f)(X)=\nabla_X\nabla f$ for $X\in
\Gamma(TM)$, and finally $V\in \Gamma(TM)$. In this article, we will
deal with the semi-elliptic case, i.e. $A$ is positive semi-definite
at each point, and we always assume that
\begin{equation}\label{222}
\sup_MTr(A)+\sup_M|V|< \infty.
\end{equation}
The purpose of this paper is to show that such a operator $L$ shares
important properties with the Laplace-Beltrami operator $\Delta$,
particularly Omori-Yau almost maximum principle and Liouville-type
theorems for subharmonic functions.
To state our main theorem, we need the following definitions.
\begin{definition}
Let $u$ be a real-valued continuous function on $M$ and let a point
$p \in M$.
\begin{itemize}
\item a function $u$ is called proper, if the set $\{p:u(p)\leq r\}$ is compact for
every real number $r$.
\item a function $v$ defined on a neighborhood $U_{p}$
of p is called an upper-supporting function for $u$ at p, if the
conditions $v(p)=u(p)$ and $v \geq u$ hold in $U_{p}$.
\end{itemize}
\end{definition}
\begin{definition}
A proper continuous function $u: M \rightarrow \mathbb{R}$ is called
an $L$-tamed exhaustion, if the following condition
holds:
\begin{enumerate}
\item $u \geq 0$.
\item At all points $p \in M$ it has a $C^2$-smooth,
upper-supporting function $v$ at $p$ defined on an open neighborhood
$U_{p}$ such that $|\nabla v|_{p} |\leq 1 $ and $L
v|_{p} \leq 1 $.
\end{enumerate}
\end{definition}
Once there is an $L$-tamed exhaustion, it easily follows that another $L$-tamed exhaustion can be chosen so that its local upper-supporting functions $v$ satisfy $|\nabla v| \leq 1 $ and $Lv \leq 1 $ not just at one point $p$ but also on its whole $U_p$.\footnote{For example, given an $L$-tamed exhaustion $u$, one can just take $\frac{u}{2}$. Then at each point $p$, $\frac{v}{2}$ can be used for an upper-supporting function which satisfies that $|\nabla \frac{v}{2}|_p| \leq \frac{1}{2} $ and $L(\frac{v}{2})|_p \leq \frac{1}{2}$. Thus by taking $U_p$ smaller (if necessary), one can achieve $|\nabla \frac{v}{2}| \leq 1$ and $L(\frac{v}{2}) \leq 1$ on each $U_p$.}
The existence of an $L$-tamed exhaustion function on a complete
Riemannian manifold is guaranteed if certain curvature conditions are satisfied. For instance,
\begin{theorem}(H.L. Royden \cite[Proposition 2]{Royden})
Every complete Riemannian manifold with its sectional curvature bounded below admits an $L$-tamed exhaustion function.
\end{theorem}
When $L$ is the Laplace-Beltrami operator $\Delta$, a stronger result holds.
K.-T. Kim and H. Lee \cite{ktk07} have shown that a $\Delta$-tamed exhaustion function exists if the Ricci
curvature $\textrm{Ric}$ satisfies
\begin{eqnarray}\label{bless}
Ric(\nabla r, \nabla r)\geq -B\rho(r)
\end{eqnarray}
for some constant $B>0$, where $r$ is the distance from an arbitrarily
fixed point in $M$ and a smooth nondecreasing function $\rho(r)$ on $[0,\infty)$ satisfies
\begin{eqnarray}\label{you}
\rho(0)=1,\ \ \ \ \int_0^\infty\frac{1}{\sqrt{\rho(t)}}\ dt=\infty,
\end{eqnarray}
\begin{eqnarray}\label{happy}
\rho^{(2k+1)}(0)=0 \ \ \forall k\geq 0,\ \ \ \ \limsup_{t\rightarrow \infty} \frac{t\rho(\sqrt{t})}{\rho(t)} < \infty.
\end{eqnarray}
For example, if
$$\textrm{Ric}(\nabla r,\nabla r) \geq -B\ r^2(\log
r)^2(\log(\log r))^2\cdots (\log^{k}r)^2$$ for $r\gg 1$, a $\Delta$-tamed exhaustion always exists.
A. Ratto, M. Rigoli, and A. Setti \cite{rrs} showed that if the above Ricci curvature condition (\ref{bless}) holds, then
for every real-valued $C^2$ function $f$ on $M$ which is bounded above,
there exists a sequence $\{p_{k}\}$ on $M$ such that
$$\lim_{k\rightarrow \infty}|\nabla f(p_{k})|=0,\
\limsup_{k\rightarrow \infty}{\Delta}f(p_{k})\leq 0, ~\textrm{and}~
\lim_{k\rightarrow \infty}f(p_{k})=\sup_{M}f.$$ This property is the
well-known Omori-Yau almost maximum principle for the Laplacian,
which was first proven by H. Omori \cite{omori} and S.T. Yau
\cite{Yau} when the Ricci curvature is only bounded below. K.-T. Kim
and H. Lee \cite{ktk07} showed the above maximum principle holds
whenever there exists a $\Delta$-tamed exhaustion. We will prove the analogous
maximum principle for the above semi-elliptic
operator $L$ also holds whenever there exists an
$L$-tamed exhaustion by following their method in
\cite{ktk07}.
L.J. Alias, D. Impera, and M. Rigoli also proved a generalized
Omori-Yau maximum principle for $L$, when the sectional
curvature $K$ satisfies
\begin{eqnarray}\label{AIR}
K(\Sigma)\geq -B\rho(r)
\end{eqnarray}
for any tangent 2-plane $\Sigma$ containing $\nabla r$, where $B>0$ is a constant, $r$ is the distance from an arbitrarily
fixed point in $M$, and $\rho(r)$ is as in (\ref{you}, \ref{happy}). (For a proof, see
\cite[Corollary 3]{alias} which is actually stated for $L$ with no first order terms but can be trivially extended to the general $L$.) It remains as a natural question whether the condition (\ref{AIR}) implies the existence of an $L$-tamed exhaustion.
Recently A. Borb\'ely \cite{bor} proved that the
Omori-Yau maximum principle for $\Delta$ holds without (\ref{happy}) in
Ratto-Rigoli-Setti's condition.
The relation between A. Borb\'ely's condition and the existence of
$\Delta$-tamed exhaustion also remains for further
study.
Now come applications. One of main applications of the Omori-Yau maximum principle is a
generalized Liouville-type theorem which gives a condition for the a
priori boundedness of solutions of Laplace-type differential
inequalities. This idea has originated from Cheng and Yau \cite{CY},
and been further extended by \cite{rrs}, \cite{suh}, \cite{cysung11},
etc. We can now extend the results of \cite{cysung11} to our semi-elliptic operator
$L$.
\begin{theorem}\label{main1}
Let $M$ be a smooth complete Riemannian manifold admitting an
$L$-tamed exhaustion function. Suppose that a $C^2$
function $f : M \rightarrow \mathbb{R}$ is bounded below and
satisfies $Lf \geq F(f)+ H(|\nabla f|)$
for real-valued continuous functions $F$ and $H$ on $\Bbb R$ such that $H(0)=0$.
\begin{enumerate}
\item If $\liminf_{x \rightarrow \infty}\frac{F(x)}{x^{\nu}}>0$ for
some $\nu>1,$ then $f$ is bounded such that $F(\sup f)\leq 0.$
\item If $\liminf_{x \rightarrow \infty}\frac{F(x)}{x^\nu}\leq 0$ for
any $\nu>1,$ then $\sup f= \infty $ or $f$ is bounded such that
$F(\sup f)\leq 0.$
\end{enumerate}
\end{theorem}
\begin{theorem}\label{main2}
Let $M$ be as in Theorem \ref{main1}. Suppose that a $C^2$ function
$f : M \rightarrow \mathbb{R}$ is bounded above and satisfies
$Lf \geq F(f)+ H(|\nabla f|)$ for $F$ and $H$ as in
the above theorem.
\begin{enumerate}
\item If $\liminf_{x \rightarrow -\infty}\frac{F(x)}{(-x)^{\nu}}>0$ for
some $\nu \leq 1,$ then $f$ is bounded such that $F(\inf f)\leq 0.$
\item If $\liminf_{x \rightarrow -\infty}\frac{F(x)}{(-x)^{\nu}}\leq 0$ for
any $\nu \leq 1,$ then $\inf f= -\infty $ or $f$ is bounded such that
$F(\inf f)\leq 0.$
\end{enumerate}
\end{theorem}
As a corollary, we give a semi-elliptic generalization of Liouville's theorem stating that any $f \in C^{2}(\mathbb{R}^{2})$ which is subharmonic $(\Delta f \geq 0)$ and bounded above must be constant.
\begin{corollary}\label{cor3}
Let $M$ be as in Theorem \ref{main1}.
\begin{enumerate}
\item There exits no $f \in C^{2}(M)$ which is bounded above and
$L f \geq c$ for a constant $c>0$.
\item Any $f \in C^{2}(M)$ which is non-positive and satisfies
$L f \geq c|f|^d$ for some positive constants $c$
and $d$ must be identically zero.
\end{enumerate}
\end{corollary}
\begin{remark}
Theorem \ref{main1} and \ref{main2} can be easily extended to any linear second-order semi-elliptic operator $$
+h,$$ for $h\in C^\infty(M)$ just by considering $Lf\geq F(f)-hf+ H(|\nabla f|).$
\end{remark}
There are many other conditions under which the Omori-Yau maximum
principle holds, and also lots of Liouville-type
theorems for a variety of subharmonic functions. For instance, the
readers may be referred to \cite{karp1, karp, Leung, nad, prs2,
prs1, prs, take, yibing, xu}, and references therein.
\section{Generalized Omori-Yau maximum principle}
\begin{theorem}\label{main3}
Let $M$ be a smooth complete Riemannian $n$-manifold admitting an
$L$-tamed exhaustion function. Then for every
real-valued $C^2$ function $f$ on $M$ which is bounded above, there
exists a sequence $\{p_{k}\}$ on $M$ satisfying the following
properties:
$$\lim_{k\rightarrow \infty}|\nabla f(p_{k})|=0,\
\limsup_{k\rightarrow \infty}Lf(p_{k})\leq 0,
~\textrm{and}~ \lim_{k\rightarrow \infty}f(p_{k})=\sup_{M}f.$$
\end{theorem}
\begin{proof}
The proof is similar to the method in the article \cite{ktk07}, which uses a sequence of compact-supported approximations of $f$, which obviously attain their maximums. Without loss of generality, we may assume that $\sup_{M}f>0$ by adding some
positive constant. Take an $L$-tamed exhaustion function $u$.
Now, we choose a point $p \in M$ such that $f(p)>0$. For each
$\epsilon>0$, let $$X_{\epsilon}=\{x \in M |
u(x)<\frac{1}{\epsilon}\}.$$ Then $X_{\epsilon}$ forms an increasing
sequence of open subsets of $M$ and each closure
$\overline{X}_{\epsilon}$ gives rise to a compact exhaustion of $M$
as $\epsilon \downarrow 0.$
Taking a positive constant $r$ such that $p \in X_{r}$. The
continuous function $$(1-ru(x))f(x)$$ vanishes on the boundary of
$X_{r}$. Since $(1-ru(p))f(p)>0$, and since $\overline{X}_{r}$ is
compact, the function $(1-ru)f$ attains its maximum value in the set
$X_{r}$, say at $p_{r} \in X_{r}$, respectively. It is obvious that
the maximum value is positive. From now on, we fix $r$.
Let $\epsilon$ be any positive constant smaller than $r$. Then $p
\in X_{r} \subset X_{\epsilon} $ and $$(1-\epsilon u(p))f(p)\geq
(1-ru(p))f(p)>0.$$ In the same way, the function $(1-\epsilon u)f$
attains a positive maximum value in the set $X_{\epsilon}$, say at
$p_{\epsilon} \in X_{\epsilon}.$
Since $A$, in the notation \eqref{eq111}, is symmetric, it is
diagonalizable at each point in an orthonormal basis, so we can take
a normal coordinate $(x_{1}, \cdots, x_{n})$ around $p_{\epsilon}
\in M$ such that $A$ at $p_{\epsilon}$ is represented as a diagonal
matrix, and hence
\begin{equation}\label{eq1}
L h|_{p_{\epsilon}}
=\sum_{l}a_{ll}(p_{\epsilon})\frac{\partial^2}{\partial
x_{l}^2}h|_{p_{\epsilon}}+\sum_{l}
a_{l}(p_{\epsilon})\frac{\partial}{\partial x_{l}}h
|_{p_{\epsilon}},
\end{equation}
for a real-valued function $h$ on $M$, where each
$a_{ll}(p_{\epsilon})$ is nonnegative, and the entries
$a_{ll}(p_{\epsilon})$ and $|a_{l}(p_{\epsilon})|$ are bounded above
as $p_{\epsilon}$ varies by \eqref{222}. For a notational
convenience, let's introduce locally-defined differential operators
\begin{equation}\label{eq10}
\widetilde{\nabla}:=(a_{11}(p_{\epsilon})\frac{\partial}{\partial
x_{1}}, ~\cdots~, a_{nn}(p_{\epsilon})\frac{\partial}{\partial
x_{n}})~~\textrm{and}~~
\widetilde{\nabla}_{1}:=a_{1}(p_{\epsilon})\frac{\partial}{\partial
x_{1}}+ ~\cdots~+ a_{n}(p_{\epsilon})\frac{\partial}{\partial
x_{n}},
\end{equation}
and put $d_{l}=a_{ll}(p_{\epsilon})$ and
$e_{l}=\sqrt{n}|a_{l}(p_{\epsilon})|$ for $1 \leq l \leq n$.
If $h$ has an extremal value at point $p_{\epsilon}$,
$$Lh |_{p_{\epsilon}}=\sum_{l}a_{ll}(p_{\epsilon})\frac{\partial^2}{\partial x_{l}^2}h|_{p_{\epsilon}}.$$
Furthermore, if a real-valued function AB on $M$ has an extremal
value at $p_{\epsilon}$, then one can obtain
\begin{equation}\label{eq2}
L(AB)|_{p_{\epsilon}}=(LA|_{p_{\epsilon}}-
\widetilde{\nabla}_{1}A|_{p_{\epsilon}})B(p_{\epsilon})+2\widetilde{\nabla}A|_{p_{\epsilon}}
\cdot \nabla B|_{p_{\epsilon}}
+A(p_{\epsilon})(LB|_{p_{\epsilon}}-
\widetilde{\nabla}_{1}B|_{p_{\epsilon}}).
\end{equation}
Note that $\widetilde{\nabla}A|_{p_{\epsilon}} \cdot \nabla
B|_{p_{\epsilon}}= \nabla A|_{p_{\epsilon}} \cdot
\widetilde{\nabla} B|_{p_{\epsilon}}.$
We may assume that $d_{1}$ and $e_{1}$ are the largest of $\{d_{1},
\cdots ,d_{n}\}$ and $\{ e_{1} , \cdots , e_{n}\}$ respectively.
Consider a $C^2$ upper-supporting function $v : U \rightarrow
\mathbb{R}$ for $u$ at $p_{\epsilon}$, where $U$ is an open
neighborhood of $p_{\epsilon}.$ Then we get $$ |\widetilde{\nabla}
v|_{p_{\epsilon}} |\leq d_{1},~ |\widetilde{\nabla}_{1}
v|_{p_{\epsilon}} |\leq e_{1},~ \textrm{and}~ L
v|_{p_{\epsilon}} \leq 1.$$ By taking $U$ further small, we may
assume that $U \subset X_{\epsilon}$ and $f$ is positive on $U$,
since $f(p_\epsilon)>0$. For every $x \in U$,
$$(1-\epsilon v(x))f(x) \leq (1-\epsilon u(x))f(x) \leq (1-\epsilon
u (p_{\epsilon}))f(p_{\epsilon})=(1-\epsilon
v(p_{\epsilon}))f(p_{\epsilon}).$$ Since $p_{\epsilon}$ is a local
maximum point of $(1-\epsilon v)f$, we get
$$\nabla [(1-\epsilon v)f]|_{p_{\epsilon}}= \widetilde{\nabla} [(1-\epsilon v)f]|_{p_{\epsilon}}=
\widetilde{\nabla}_{1}[(1-\epsilon v)f]|_{p_{\epsilon}}=0 .$$ By a
simple calculation, we have
$$ (1-\epsilon v(p_{\epsilon}))| \widetilde{\nabla} f(p_{\epsilon}) |=
\epsilon | \widetilde{\nabla} v(p_{\epsilon}) | f(p_{\epsilon})
\leq \epsilon d_{1}( \sup_{M}f).$$ From
$v(p_{\epsilon})=u(p_{\epsilon})$, we get
$$(1-\epsilon u(p_{\epsilon}))| \widetilde{\nabla} f(p_{\epsilon}) | \leq \epsilon d_{1}(\sup_{M}f).$$
Also, because $ X_{r} \subset X_{\epsilon}$, we have
$$(1-\epsilon u(p_{r}))f(p_{r}) \leq (1-\epsilon
u(p_{\epsilon}))f(p_{\epsilon}).$$ This implies that
\begin{eqnarray*}
(1-ru(p_{r}))f(p_{r})|\widetilde{\nabla}f(p_{\epsilon})|&\leq&(1-\epsilon
u(p_{r}))f(p_{r})|\widetilde{\nabla}f(p_{\epsilon}) | \leq
(1-\epsilon
u(p_{\epsilon}))f(p_{\epsilon})|\widetilde{\nabla}f(p_{\epsilon})
|\\&\leq& f(p_{\epsilon})\epsilon d_{1} (\sup_{M}f) \leq \epsilon
d_{1}(\sup_{M}f)^{2}.
\end{eqnarray*}
So, we conclude that
$$ |\widetilde{\nabla}f(p_{\epsilon}) | \leq \epsilon \frac{
d_{1}(\sup_{M}f)^{2}}{(1-r u(p_{r})) f(p_{r})}.$$ Note that
$K:=\frac{d_{1}(\sup_{M}f)^{2}}{(1-ru(p_{r})) f(p_{r})}$ is a
positive constant independent of $\epsilon$ with $\epsilon<r$.
Therefore, we obtain $$\lim_{\epsilon\rightarrow
0}|\widetilde{\nabla} f (p_{\epsilon})|=0.$$ By the same method as
above, we have
$$ | \nabla f(p_{\epsilon}) | \leq \epsilon \frac{
(\sup_{M}f)^{2}}{(1-r u(p_{r})) f(p_{r})} \ \ \ \ \textrm{and}\ \ \ \
|\widetilde{\nabla}_{1}f(p_{\epsilon}) | \leq \epsilon \frac{
e_{1}(\sup_{M}f)^{2}}{(1-r u(p_{r})) f(p_{r})}$$ Therefore, we get
$$\lim_{\epsilon \rightarrow 0}|\nabla f (p_{\epsilon})|=0\ \ \ \
\textrm{and}\ \ \ \ \lim_{\epsilon \rightarrow
0}|\widetilde{\nabla}_{1} f (p_{\epsilon})|=0.$$ Now we prove
$$\limsup_{\epsilon \rightarrow
0}Lf(p_{\epsilon})\leq 0.$$ Since $p_{\epsilon}$ is
a local maximum point of $(1-\epsilon v)f$, we have
$L((1-\epsilon v )f) \leq 0$ at point
$p_{\epsilon}.$ Using the formula (\ref{eq2}),
\begin{eqnarray*}
[L((1-\epsilon v )f)]|_{p_{\epsilon}} &=&-\epsilon
Lv|_{p_{\epsilon}}f(p_{\epsilon})+\epsilon
\widetilde{\nabla}_{1} v|_{p_{\epsilon}}f(p_{\epsilon})-2\epsilon
\nabla v|_{p_{\epsilon}} \cdot \widetilde{\nabla} f|_{p_{\epsilon}}+
(1-\epsilon v(p_{\epsilon}))Lf|_{p_{\epsilon}}
\\
& &- (1-\epsilon v(p_{\epsilon}))
\widetilde{\nabla}_{1}f|_{p_{\epsilon}}\\ &\leq& 0.
\end{eqnarray*}
Hence
\begin{eqnarray*}
(1-\epsilon v(p_{\epsilon}))Lf|_{p_{\epsilon}}
&\leq& 2\epsilon \nabla v|_{p_{\epsilon}} \cdot \widetilde{\nabla}
f|_{p_{\epsilon}} +\epsilon
Lv|_{p_{\epsilon}}f(p_{\epsilon})-\epsilon
\widetilde{\nabla}_{1} v|_{p_{\epsilon}}f(p_{\epsilon})+(1-\epsilon
v(p_{\epsilon}))
\widetilde{\nabla}_{1}f|_{p_{\epsilon}}\\
&\leq& \epsilon (2|\widetilde{\nabla }f|_{p_{\epsilon}} |+
\sup_{M}f+ e_{1}\sup_{M}f)+|\widetilde{\nabla}_{1}f|_{p_{\epsilon}}|
(1-\epsilon v(p_{\epsilon}))\\
&\leq& \epsilon ( 2\epsilon K + \sup_{M}f+ e_{1}\sup_{M}f
)+|\widetilde{\nabla}_{1}f|_{p_{\epsilon}}| (1-\epsilon
v(p_{\epsilon})).
\end{eqnarray*}
Since $1-\epsilon u(p_{\epsilon})=1-\epsilon v(p_{\epsilon})>0,$
we get
$$Lf|_{p_{\epsilon}}\leq \epsilon \frac {(2
\epsilon K + \sup_{M}f + e_{1}\sup_{M}f)} {(1-\epsilon
u(p_{\epsilon}))}+|\widetilde{\nabla}_{1}f|_{p_{\epsilon}}|.$$ As
above, we obtain
\begin{eqnarray*}
Lf|_{p_{\epsilon}}&\leq& \epsilon \frac {(2
\epsilon K + \sup_{M}f + e_{1}\sup_{M}f)(\sup_{M}f)} {(1-\epsilon
u(p_{\epsilon}))f(p_{\epsilon})}+\epsilon \frac{
e_{1}(\sup_{M}f)^{2}}{(1-r u(p_{r})) f(p_{r})}\\ &\leq& \epsilon
\frac {(2 \epsilon K + \sup_{M}f + e_{1}\sup_{M}f)(\sup_{M}f)} {(1-r
u(p_{r}))f(p_{r})}+\epsilon \frac{ e_{1}(\sup_{M}f)^{2}}{(1-r
u(p_{r})) f(p_{r})}.
\end{eqnarray*}
Therefore, we conclude that there is a positive constant $C$
independent of $\epsilon$ such that $Lf|_{p_{\epsilon}} \leq C \epsilon.$
It only remains to show that $\lim_{\epsilon\rightarrow
0}f(p_{\epsilon})=\sup_{M} f.$
Let $\eta$ be any positive constant such that $\sup_{M} f>\eta $. We
may choose a point $q \in M$ such that
$f(q)>\sup_{M}f-\frac{\eta}{2}$. Also choosing a positive constant
$\epsilon$ with $\epsilon < r$ such that $q\in X_{\epsilon}$ and
$\epsilon u(q) f(q) \leq \frac{\eta}{2}$ , we get
$$(1-\epsilon u(p_{\epsilon}))f(p_{\epsilon}) \geq
(1-\epsilon u(q))f(q) \geq \sup_{M}f-\eta.$$ Since $0<1-\epsilon
u(p_{\epsilon})<1,$ we have
$$f(p_{\epsilon})\geq
\frac{\sup_{M}f-\eta}{1-\epsilon u(p_{\epsilon})} >
\sup_{M}f-\eta,$$ completing the proof.
\end{proof}
\begin{remark}
L.J. Alias, D. Impera, and M. Rigoli used their generalized
Omori-Yau maximum principle to obtain certain estimates
of higher order mean curvatures of hypersurfaces in some warped
product spaces, and D. Impera \cite{DI} similarly obtained such
estimates for spacelike hypersurfaces in Lorentzian manifolds.
\end{remark}
\section{Proof of Theorem \ref{main1}}
We follow the idea of \cite{suh, cysung11}. We may choose a constant $a$
such that $f+a>0,$ because $f$ is bounded below. Let $G:M
\rightarrow \mathbb{R}^{+}$ be a $C^2$ function defined by
$G=(f+a)^{\frac{1-q}{2}}$ where $q>1$ is a constant.
Since $G$ is bounded below, Theorem \ref{main3} implies that for
any $\delta >0$ there exists a point $p_{\epsilon} \in M$ such that
\begin{equation}\label{eq3}
|\nabla G(p_{\epsilon})|<\delta,~ |\widetilde{\nabla}
G(p_{\epsilon})| <\delta,~
LG(p_{\epsilon})>-\delta, ~ \textrm{and} ~ \inf
G+\delta>G(p_{\epsilon}),
\end{equation} where $\widetilde{\nabla}$ is defined by
(\ref{eq10}). Note that $G(p_{\epsilon}) \rightarrow \inf G$ and
$f(p_{\epsilon}) \rightarrow \sup f$ as $\delta \rightarrow 0.$
By a direct calculation,
\begin{equation}\label{eq4}
\widetilde{\nabla}G|_{p_{\epsilon}}=
(\frac{1-q}{2})G(p_{\epsilon})^{\frac{q+1}{q-1}}\widetilde{\nabla}
f|_{p_{\epsilon}}.
\end{equation}
\begin{lemma}
\begin{equation}\label{eq5}
LG|_{p_{\epsilon}}=
-(\frac{q+1}{2})G(p_{\epsilon})^{\frac{2}{q-1}}\nabla
G|_{p_{\epsilon}} \cdot \widetilde{\nabla}
f|_{p_{\epsilon}}+(\frac{1-q}{2})G(p_{\epsilon})^{\frac{q+1}{q-1}}Lf|_{p_{\epsilon}}.
\end{equation}
\end{lemma}
\begin{proof}
By (\ref{eq1}), evaluating $LG$ at $p_{\epsilon}$,
we have
$$LG|_{p_{\epsilon}}
=\sum_{l}a_{ll}(p_{\epsilon})\frac{\partial^2}{\partial
x_{l}^2}G|_{p_{\epsilon}}+\sum_{l}
a_{l}(p_{\epsilon})\frac{\partial}{\partial x_{l}}G
|_{p_{\epsilon}},~ \textrm{where}~ 1 \leq l \leq n.$$ By a simple
calculation, one gets
$$\sum_{l}a_{ll}(p_{\epsilon})\frac{\partial^2}{\partial x_{l}^2}G|_{p_{\epsilon}}=-(\frac{q+1}{2})G(p_{\epsilon})^{\frac{2}{q-1}}\nabla
G|_{p_{\epsilon}} \cdot \widetilde{\nabla} f|_{p_{\epsilon}}+
\sum_{l}a_{ll}(p_{\epsilon})(\frac{1-q}{2})(f(p_{\epsilon})+a)^{\frac{-1-q}{2}}\frac{\partial^2}{\partial
x_{l}^2}f|_{p_{\epsilon}}$$ and $$\sum_{l}
a_{l}(p_{\epsilon})\frac{\partial}{\partial x_{l}}G
|_{p_{\epsilon}}=\sum_{l}
a_{l}(p_{\epsilon})(\frac{1-q}{2})(f(p_{\epsilon})+a)^{\frac{-1-q}{2}}\frac{\partial}{\partial
x_{l}}f|_{p_{\epsilon}}.$$ This yields the desired equality.
\end{proof}
By plugging (\ref{eq4}) to (\ref{eq5}), we have
$$(\frac{1-q}{2})G(p_{\epsilon})^{\frac{2q}{q-1}}Lf|_{p_{\epsilon}}=G(p_{\epsilon})LG|_{p_{\epsilon}}-(\frac{q+1}{q-1})\nabla
G(p_{\epsilon}) \cdot \widetilde{\nabla} G(p_{\epsilon}).$$ Applying
(\ref{eq3}) gives
\begin{equation}\label{eq6}
(\frac{1-q}{2})G(p_{\epsilon})^{\frac{2q}{q-1}}Lf|_{p_{\epsilon}}>
G(p_{\epsilon})(-\delta)-(\frac{q+1}{q-1})\delta^{2}.
\end{equation}
Applying $Lf \geq F(f)+ H(|\nabla f|)$ and
replacing $G$ by $(f+a)^{\frac{1-q}{2}},$ we have
\begin{equation}\label{eq7}
\frac{F(f(p_{\epsilon}))+H(|\nabla
f(p_{\epsilon})|)}{(f(p_{\epsilon})+a)^{q}}<(\frac{2\delta}{q-1})
\frac{1}{(f(p_{\epsilon})+a)^{\frac{q-1}{2}}}+\frac{2(q+1)}{(q-1)^{2}}\delta^{2}.
\end{equation}
Assume that $\sup f< \infty.$ Then as $\delta \rightarrow 0,$ since $\nabla G|_{p_{\epsilon}}\rightarrow 0,$ $G$ is bounded below by a positive constant, and
$$\nabla G|_{p_{\epsilon}}=
(\frac{1-q}{2})G(p_{\epsilon})^{\frac{q+1}{q-1}}\nabla
f|_{p_{\epsilon}},$$ we have $H(|\nabla f(p_{\epsilon})|)\rightarrow
0.$ Also, the $\mathbf{RHS}$ of (\ref{eq7}) converges to $0$ while
the $\mathbf{LHS}$ of (\ref{eq7}) converges to $\frac{F(\sup
f)}{(\sup f + a)^{q}}$ as $\delta \rightarrow 0.$ Thus, we get
$F(\sup f)\leq 0.$
Finally, it remains to show that when $\liminf_{x\rightarrow
\infty}\frac{F(x)}{x^{\nu}}>0$ for some $\nu >1,$ $f$ must be
bounded. Assume to the contrary that $\sup f = \infty.$ Then for $q< \nu,$ the
$\mathbf{RHS}$ of (\ref{eq7}) converges to $0$, while the $\mathbf{LHS}$ of (\ref{eq7}) diverges to
$\infty$ as $\delta
\rightarrow 0.$ This is a desired contradiction, which completes the proof.
\section{Proof of Theorem \ref{main2}}
We again follow the idea of \cite{suh, cysung11}. Since $-f$ is bounded below,
we can apply the proof of Theorem \ref{main1} to $-f$ with $q<1$. By
the inequality (\ref{eq6}), we get
$$(\frac{1-q}{2})G(p_{\epsilon})^{\frac{2q}{q-1}}L(-f)|_{p_{\epsilon}}>
G(p_{\epsilon})(-\delta)-\frac{|q+1|}{|q-1|}\delta^{2}.$$ Applying
$Lf \geq F(f)+ H(|\nabla f|),$ we have
$$\frac{F(f(p_{\epsilon}))+H(|\nabla f(p_{\epsilon})|)}{(-f(p_{\epsilon})+a)^{q}}\leq
\frac{L
f(p_{\epsilon})}{(-f(p_{\epsilon})+a)^{q}}<(\frac{2\delta}{1-q})
\frac{1}{(-f(p_{\epsilon})+a)^{\frac{q-1}{2}}}+\frac{2|q+1|}{(q-1)^{2}}\delta^{2}.$$
By a simple calculation,
\begin{equation}\label{eq8}
\frac{F(f(p_{\epsilon}))+H(|\nabla
f(p_{\epsilon})|)}{(-f(p_{\epsilon})+a)^{\frac{q+1}{2}}}<
\frac{2\delta}{1-q}+\frac{2|q+1|}{(q-1)^{2}}\delta^{2}(-f(p_{\epsilon})+a)^{\frac{q-1}{2}}.
\end{equation}
By the same method as above, we get $H(|\nabla
f(p_{\epsilon})|)\rightarrow 0.$ If $\inf f >- \infty,$ then $F(\inf
f) \leq 0 $ as $\delta \rightarrow 0.$
Now it only remains to show that if $\liminf_{x\rightarrow
-\infty}\frac{F(x)}{(-x)^{\nu}}>0$ for some $\nu \leq 1,$ then $f$ is
bounded. Let's assume that to the contrary $\inf f=- \infty.$ By taking $q$ such that
$\frac{q+1}{2}< \nu$ and letting $\delta \rightarrow 0$, the
$\mathbf{RHS}$ of (\ref{eq8}) converges to $0$ while the
$\mathbf{LHS}$ of (\ref{eq8}) diverges to $\infty$. This is a
contradiction completing the proof.
\section{Proof of Corollary \ref{cor3}}
Suppose that $f$ is bounded above and satisfies $L f \geq c>0$ for a
constant $c$. Applying Theorem \ref{main2} with $F=c$ and $H=0$, one conclude
that $f$ is bounded and $F(\inf f) \leq 0.$ This is
contradictory to $F\equiv c
>0.$
For a proof of Corollary \ref{cor3} $(2)$, applying Theorem
\ref{main2} with $F(f)=c|f|^{d},$ it follows that $f$ is bounded and
$c|\inf f|^{d} \leq 0$ implying $f \equiv 0.$
\section*{Acknowledgments}
The authors would like to thank Hanjin Lee for remarks about the tamed exhaustion function.
This work was supported by the National Research Foundation of Korea(NRF) grant
funded by the Korea government(MEST) (No. 2012-0000341, 2011-0002791).
\footnotesize
\bibliographystyle{amsplain}
\providecommand{\bysame}{\leavevmode\hbox
to3em{\hrulefill}\thinspace}
|
\section{Introduction}
Celestial Mechanics systems have two fundamental conservation principles that enable their deeper analysis: conservation of momentum and conservation of (mechanical) energy. Of the two, conservation of momentum provides the most constraints on a general system, with three translational symmetries (which can be trivially removed) and three rotational symmetries. If no external force acts on the system, these quantities are always conserved independent of the internal interactions of the system. Conservation of energy instead involves assumptions on both the lack of exogenous forces and on the nature of internal interactions within the system. For this reason energy is often not conserved for ``real'' systems that involve internal interactions, such as tidal deformations or impacts, even though they may conserve their total momentum. Thus mechanical energy generally decays through dissipation until the system has found a local or global minimum energy configuration that corresponds to its constant level of angular momentum.
This observation motivates a fundamental question for celestial mechanics: \\
{\centering\it What is the minimum energy configuration of a $N$-body system with a fixed level of angular momentum?}\\
This paper shows that this is an ill-defined question for traditional point-mass celestial mechanics systems. If instead the system and problem are formulated accounting for finite density distributions this question becomes well posed and provides new light on celestial mechanics systems.
This is a well motivated adjustment as real systems always have a finite density and, hence, any particle in a celestial mechanics system has a finite size. Such a physically corrected system has been called the ``Full $N$-Body Problem,'' as inclusion of finite density also necessitates the modeling of the rotational motion of the components, which is not needed for consideration of point masses. It also necessitates consideration of contact forces as their mass centers cannot come arbitrarily close to each other, as at some distance they will rest on each other. Thus, introduction of this finite density correction allows the minimum energy configurations for an $N$-body system to be explicitly defined and computed for a given level of angular momentum.
Given this perspective, an interesting problem is to track the absolute minimum energy configuration of a collection of $N$ particles as the system angular momentum increases from or decreases to zero. This is, essentially, an investigation of the celestial mechanics of granular systems as a function of total angular momentum. This problem has been shown to be relevant to the understanding of solar system bodies, especially among asteroids whose size is small enough so that when the components rest on each other they have insufficient gravitational attraction to overcome material strength, and thus retain the physical characteristics of rigid bodies resting on each other. Due to the celebrated YORP effect, which has established that non-symmetrically shaped bodies subject to solar radiation will change their spin rates over time, this question has several practical applications and has been implicated in how small asteroids form binary systems \cite{scheeres_fission}.
The question of stable minimum energy configurations for the $N=2$ particle Full Body problem has been worked out in detail \cite{scheeres_gravgrad, scheeres_F2BP_planar}, and has been verified as a viable physical model via astronomical observations of asteroids \cite{pravec_fission}.
This paper presents a number of Theorems that motivate and enable our study of minimum energy configurations, and then explore some exact results for the cases of $N=1,2,3$. Some hypotheses for the case $N\gg 1$ are also given.
\begin{description}
\item[Theorem 1:] A sharper version of the Sundman Inequality is derived: $H^2 \le 2 I_H T \le 2 I_p T$, where $H$ is the system angular momentum, $T$ is the system kinetic energy, $I_p$ is the system polar moment of inertia and $I_H$ is the system moment of inertia about a fixed direction in space. This form of the Sundman Inequality is the appropriate generalization for bodies with finite density distributions.
\item[Theorem 2:] The minimum energy function ${\cal E}$ is defined: ${\cal E} = \frac{H^2}{2I_H} + U \le E$, where $U$ is the total potential energy of the system and $E$ is the total system energy. It is shown that ${\cal E} = E$ occurs if and only if the minimum energy function is stationary with respect to all variations for a given level of angular momentum $H$ and represents a relative equilibrium for the system (borrowing from a theorem originally proven by Smale).
\item[Theorem 3:] It is shown that the point-mass $N$-body problem with a specified level of angular momentum does not have either global or local minimum energy configurations for all $N\ge 3$ (borrowing from a theorem by Moeckel).
\item[Theorem 4:] It is shown that the finite-density $N$-body problem has a global minimum energy configuration for all $N\ge 1$ and for all levels of angular momentum.
\item[Theorem 5-8:] All minimum energy configurations, both local and global, are found for all levels of angular momentum for the finite density cases $N=1,2,3$.
\item[Hypothesis 1:] It is hypothesized that for all $N$ the local and global minimum energy configurations for a finite density system will either condense into a single aggregate and spin at a uniform rate, or will lie in a two component system with the orbit and all rotations spinning at the same uniform rate.
\end{description}
The current paper proves a number of fundamental results that motivate and enable the further exploration of these questions.
These results show a surprising complexity in the evolution of minimum energy states as a function of angular momentum, with distinctly different pathways arising as the number of particles in the system increases.
\section{Background}
\subsection{Conservation Principles}
A closed mechanical system has two fundamental conservation principles: conservation of momentum, both translational and rotational, and conservation of energy. The statement of these is simple and can be derived starting from the fundamental equations of motion for this system. This paper does not detail these results, as they are well known (see the explicit derivation of these conservation principles for the full 2-body problem in \cite{scheeres_F2BP}).
The result is that the total translational momentum, angular momentum $\bfm{H}$ and total energy $E = T+U$ (where $T$ is kinetic energy and $U$ is gravitational potential energy) is conserved under motion within the system. By specifying the barycenter of the system to be at rest and at the origin of the system the three components of the translational momentum are removed, which allows the fundamental quantities of the system to be stated in a purely relative form. Angular momentum of the system is also conserved so long as no exogenous forces act -- regardless of the internal actions between the components. The same is not true of the energy, as its conservation depends on both the lack of exogenous forces and on strict specification of internal interactions. Specifically, all internal interactions must be reversible, meaning that no dissipation can occur. For ideal celestial mechanics involving point mass interactions this is a reasonable model -- however the point-mass celestial mechanics model is always an approximation due to its infinite density distribution at the particle location. The truth is that all physical celestial mechanics systems of interest involve dissipative interactions, and hence all natural systems will tend to dissipate energy, meaning that total energy is no longer conserved. In many cases the timescale for energy dissipation is quite slow, allowing for the system to be adequately modeled over extended periods of time as conserving energy. However, as has been well established in the study of the solar system, energy dissipation due to tidal distortion of bodies and internal stress wave propagation eventually dominates the dynamics and interactions of almost all celestial mechanics systems and forces them to evolve to lower energy states, all the while conserving total angular momentum \cite{goldreich_peale,goldreich}.
Two simple examples of this can be mentioned. First is that almost all minor planets in the solar system are in or close to uniform rotation about their maximum moments of inertia -- this despite a past filled with mutual impacts and gravitational torques from their occasional interactions with planets. This is simply due to the fact that uniform rotation about the maximum moment of inertia is the minimum energy state for a solo rotating body. The actual dissipation occurs due to the interaction of time varying accelerations that the body experiences when not in uniform rotation with the material properties of the body, causing minute elastic deformations that turn mechanical energy into heat, which is then radiated away \cite{burns_safronov}. The other example is the mutual orbit of the Earth and moon. The moon has already shed its excess rotational energy and has settled into a local minimum energy state, with the same face always pointing at the Earth. The Earth still spins more rapidly than the mutual orbit, however, raising tides on the Earth which dissipate energy and slow the Earth's rotation rate. Conservation of angular momentum dictates that the mutual orbit expands -- this occurs physically by the torque that the Earth's tidal bulge places on the moon. The eventual state of the Earth-Moon system (ignoring the sun) is for both systems to devolve into mutually synchronous orbits, as is the case for the Pluto-Charon system. Once in this final state no excess energy dissipation can occur as the system is in its minimum energy state for its given total angular momentum.
\subsection{Density Distributions and Fundamental Quantitites}
To discuss minimum energy configurations for the point mass or finite density $N$ body problem requires the definition of a few basic scalar and vector functions and quantities. Consider the total kinetic energy, gravitational potential energy, polar moment, and angular momentum of an arbitrary collection of $N$ mass distributions, denoted as ${\cal B}_i, i = 1,2, \ldots, N$. Each body ${\cal B}_i$ is defined by a differential mass distribution $dm_i$ that can fall into one of two forms, a point mass distribution or a finite density distribution, denoted as
\begin{eqnarray}
dm_i & = & \left\{ \begin{array}{cc}
m_i \delta(|\bfm{r}-\bfm{r}_i|) & \mbox{ Point Mass Density Distribution} \\
\rho_i(\bfm{r}) \ dV & \mbox{ Finite Density Distribution} \end{array} \right. \label{eq:dmass}
\end{eqnarray}
where $m_i$ is the total mass of body ${\cal B}_i$, $\delta(-)$ is the Dirac delta function, $|-|$ denotes the Euclidian norm, $\rho_i$ is the density of body ${\cal B}_i$ (possibly constant) and $dV$ represents the differential volume of the body. If ${\cal B}_i$ is described by a point mass density distribution, the body itself is just defined as a single point $\bfm{r}_i$. Instead, if the body is defined as a finite density distribution, ${\cal B}_i$ is defined as a compact set in $\mathbb{R}^3$ over which $\rho_i(\bfm{r}) \ne 0$. In either case the ${\cal B}_i$ are defined as compact sets.
Assume that each differential mass element $dm_i(\bfm{r})$ has a specified position and an associated velocity. For components within a given body ${\cal B}_i$ a rigid body assumption is made so that the entire body can be defined by the position and velocity of its center of mass, its attitude, and its angular velocity. Finally, assume that these positions and velocities are defined relative to the system barycenter, which is chosen as the origin, or
\begin{eqnarray}
\sum_{i=1}^N \int_{{\cal B}_i} \bfm{r} dm_i(\bfm{r}) & = & 0 \\
\sum_{i=1}^N \int_{{\cal B}_i} \dot{\bfm{r}} dm_i(\bfm{r}) & = & 0
\end{eqnarray}
Given these definitions an integral form of the kinetic energy, polar moment, gravitational potential energy, and angular momentum vector can be stated as
\begin{eqnarray}
T & = & \frac{1}{2} \sum_{i=1}^N \int_{{\cal B}_i} \left( \dot{\bfm{r}}\cdot\dot{\bfm{r}} \right) dm_i(\bfm{r}) \\
I_p & = & \sum_{i=1}^N \int_{{\cal B}_i} \left( {\bfm{r}}\cdot{\bfm{r}} \right) dm_i(\bfm{r}) \label{eq:isum} \\
U & = & - {\cal G} \sum_{i=1}^{N-1} \sum_{j=i+1}^N \int_{{\cal B}_i} \int_{{\cal B}_j} \frac{dm_i \ dm_j}{|\bfm{r}_{ij}|} \label{eq:udef0} \\
\bfm{H} & = & \sum_{i=1}^N \int_{{\cal B}_i} \left( {\bfm{r}}\times\dot{\bfm{r}} \right) dm_i(\bfm{r}) \label{eq:hsum}
\end{eqnarray}
where $\bfm{r}_{ij} = \bfm{r}_j - \bfm{r}_i$. Note that the definition of $U$ in Eqn.\ \ref{eq:udef0} eliminates the self-potentials of these bodies from consideration. As the finite density mass distributions are rigid bodies this elimination is reasonable.
This notation can be further generalized by defining the single and joint general mass differentials
\begin{eqnarray}
dm(\bfm{r}) & = & \sum_{i=1}^N dm_i(\bfm{r}) \\
dm(\bfm{r}) \ dm'(\bfm{r}') & = & \sum_{i=1}^{N-1} \sum_{j=i+1}^N dm_i(\bfm{r}) \ dm'_j(\bfm{r}')
\end{eqnarray}
and the total mass distribution ${\cal B} = \left\{ {\cal B}_i, i = 1, 2, \ldots, N\right\}$. Then the above definitions can be reduced to integrals over ${\cal B}$:
\begin{eqnarray}
T & = & \frac{1}{2} \int_{{\cal B}} \left( \dot{\bfm{r}}\cdot\dot{\bfm{r}} \right) dm(\bfm{r}) \\
I_p & = & \int_{{\cal B}} \left( {\bfm{r}}\cdot{\bfm{r}} \right) dm(\bfm{r}) \label{eq:ipint} \\
U & = & - {\cal G} \int_{{\cal B}} \int_{{\cal B}} \frac{dm \ dm'}{|\bfm{r}-\bfm{r}'|}\\
\bfm{H} & = & \int_{{\cal B}} \left( {\bfm{r}}\times\dot{\bfm{r}} \right) dm(\bfm{r}) \label{eq:hint} \\
\int_{\cal B} \bfm{r} dm(\bfm{r}) & = & 0 \\
\int_{\cal B} \dot{\bfm{r}} dm(\bfm{r}) & = & 0
\end{eqnarray}
\subsection{Point-Mass Results}
Assuming that all of the differential densities follow the point mass density distribution, the classical results for the $N$-body problem are recovered:
\begin{eqnarray}
T & = & \frac{1}{2} \sum_{i=1}^N m_i \left( \dot{\bfm{r}}_i\cdot\dot{\bfm{r}}_i \right) \\
I_p & = & \sum_{i=1}^N m_i \left( {\bfm{r}}_i\cdot{\bfm{r}}_i \right) \\
U & = & - {\cal G} \sum_{i=1}^{N-1} \sum_{j=i+1}^N \frac{m_i \ m_j}{|\bfm{r}_{ij}|}\\
\bfm{H} & = & \sum_{i=1}^N m_i \left( {\bfm{r}}_i\times\dot{\bfm{r}}_i \right)
\end{eqnarray}
It is convenient to apply Lagrange's Identity to the kinetic energy, polar moment of inertia and angular momentum, allowing them to be restated in a relative form. Note that the following relationships only hold if the barycenter of the system is at the origin.
\begin{eqnarray}
T & = & \frac{1}{2M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( \dot{\bfm{r}}_{ij}\cdot\dot{\bfm{r}}_{ij} \right) \\
I_p & = & \frac{1}{M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( {\bfm{r}}_{ij}\cdot{\bfm{r}}_{ij} \right) \\
\bfm{H} & = & \frac{1}{M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( {\bfm{r}}_{ij}\times\dot{\bfm{r}}_{ij} \right)
\end{eqnarray}
\subsection{Full-Body Results}
If finite density distributions are assumed for each body the point-mass results must be generalized to incorporate rotational kinetic energy, rigid body moments of inertia, angular velocities and explicit mutual potentials that are a function of body attitude\cite{scheeres_F2BP}. In the following the $i$th rigid body's center of mass is located by the position $\bfm{r}_i$ and has a velocity $\dot{\bfm{r}}_i$. In addition to its mass $m_i$, the $i$th body has an inertia dyadic $\bfm{I}_i$, an angular velocity vector $\bfm{\Omega}_i$ and an attitude dyadic that maps its body-fixed vectors into inertial space, $\bfm{A}_i$. The basic quantities are then defined as (generalizing results from \cite{scheeres_F2BP}):
\begin{eqnarray}
T & = & \frac{1}{2} \sum_{i=1}^N \left[ m_i \left( \dot{\bfm{r}}_i\cdot\dot{\bfm{r}}_i \right) + \bfm{\Omega}_i \cdot \bfm{I}_i \cdot \bfm{\Omega}_i \right] \\
I_p & = & \sum_{i=1}^N m_i \left[ \left( {\bfm{r}}_i\cdot{\bfm{r}}_i \right) + \frac{1}{2} \mbox{Trace}( \bfm{I}_{i} )\right] \\
U & = & \sum_{i=1}^{N-1} \sum_{j=i+1}^N U_{ij}( \bfm{r}_{ij}, \bfm{A}_{ij} ) \label{eq:udef} \\
\bfm{H} & = & \sum_{i=1}^N \left[ m_i \left( {\bfm{r}}_i\times\dot{\bfm{r}}_i \right) + \bfm{A}_i\cdot\bfm{I}_i\cdot\bfm{\Omega}_i \right]
\end{eqnarray}
In the above the inertia dyadics are all specified in a body-fixed frame and thus are constant, the $U_{ij}$ are mutual potentials between two different rigid bodies $i$ and $j$ and are only a function of their relative position and relative attitude, $\bfm{A}_{ij}$, equal to $\bfm{A}_j^T\cdot\bfm{A}_i$, and which transfers a vector from the body $i$ frame into the body $j$ frame.
The kinetic energy, polar moment of inertia and angular momentum can again be stated in relative form between the center of masses, leaving the rotational components in their current form.
\begin{eqnarray}
T & = & \frac{1}{2M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( \dot{\bfm{r}}_{ij}\cdot\dot{\bfm{r}}_{ij} \right) + \frac{1}{2} \sum_{i=1}^N \bfm{\Omega}_i \cdot \bfm{I}_i \cdot \bfm{\Omega}_i \\
I_p & = & \frac{1}{M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( {\bfm{r}}_{ij}\cdot{\bfm{r}}_{ij} \right) + \frac{1}{2} \sum_{i=1}^N \mbox{Trace}( \bfm{I}_i) \\
\bfm{H} & = & \frac{1}{M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( {\bfm{r}}_{ij}\times\dot{\bfm{r}}_{ij} \right) + \sum_{i=1}^N \bfm{A}_i\cdot\bfm{I}_i\cdot\bfm{\Omega}_i
\end{eqnarray}
\subsubsection{Alternate System Moment of Inertia}
An alternate version of the system moment of inertia can be defined once the direction of the total angular momentum is specified. First define the total inertia dyadic of the $N$ body system as
\begin{eqnarray}
\bfm{I} & = & - \int_{\cal B} \tilde{\bfm{r}}\cdot\tilde{\bfm{r}} dm
\end{eqnarray}
where $\tilde{\bfm{r}}$ is the cross product dyadic defined such that $\bfm{a}\times\bfm{b} = \tilde{\bfm{a}}\cdot\bfm{b} = \bfm{a}\cdot\tilde{\bfm{b}}$. When evaluated in detail, and in a common inertial frame, the system inertia dyadic is equal to
\begin{eqnarray}
\bfm{I} & = & \sum_{i=1}^N \left[ m_i \left( r_{i}^2 \bfm{U} - \bfm{r}_{i}\bfm{r}_{i}\right) + \bfm{A}_i\cdot\bfm{I}_i\cdot\bfm{A}_i^T \right]
\end{eqnarray}
where the identity $- \tilde{\bfm{a}}\cdot\tilde{\bfm{b}} = (\bfm{a}\cdot\bfm{b}) \bfm{U} - \bfm{b}\bfm{a}$ is applied and where $\bfm{U}$ is the unity dyadic and the multiplication of two vectors is defined as a dyad. Given a defined direction of the total angular momentum vector in inertial space, $\hat{\bfm{H}}$, the moment of inertia relative to this direction can be defined as
\begin{eqnarray}
I_H & = & \hat{\bfm{H}}\cdot\bfm{I}\cdot\hat{\bfm{H}} \label{eq:ih}
\end{eqnarray}
This alternate version of the polar moment of inertia is of use in developing sharper limits for the Sundman Inequality later.
For the point mass density distributions all of the moments of inertia disappear and $I_p = I_H$ if the bodies and their velocities all lie in a common plane.
\subsubsection{Minimum Distances between Mass Collections}
A crucial aspect of the finite density distributions are that these bodies can come into contact with each other without singularities or deformations. Thus, under a rigid body assumption they will have minimum distances between their centers of mass as a function of their relative attitude. If the bodies are convex, strong results on the minimum distances between the bodies can be found. However, even complex shaped bodies will still have well defined minimum distances between the mass collections.
Each body has a minimum diameter (equal to the diameter of the inscribing sphere and denoted by $d$) and a maximum diameter (equal to the diameter of the circumscribing sphere and denoted by $D$), these are equal only if the body is a sphere. The relative distance between any two mass distributions ${\cal B}_i$ and ${\cal B}_j$ with minimum diameters $d_i$ and $d_j$ is bounded from below by the average of these, or $r_{ij} \ge \frac{1}{2}(d_i + d_j) = d_{ij}$. For distances $r_{ij} > \frac{1}{2}(D_i + D_j) = D_{ij}$ the relative position and attitude between the two collections is unconstrained. For any combination of relative position $D_{ij}\hat{\bfm{r}}_{ij}$ and relative attitude $\bfm{A}_{ij}$ the distance between the two mass distributions can be decreased to their relative distance at which they will touch, defined as $d_{ij}(\hat{\bfm{r}}_{ij}, \bfm{A}_{ij})$. The contact points of bodies $i$ and $j$ that define this distance lie in their respective boundary sets of the mass distributions. For the given definition, the set of these minimum distances is a closed set and is bounded from below by $d_{ij}$, the minimum distance. For non-convex shapes these minimum distance sets can be complicated and may not even be able to attain the theoretical minimum. Also, in these cases different definitions of the minimum distance can be defined, depending on whether the bodies are star-convex or have a more convoluted shape. Note that across any of these shapes it is always possible to define a minimum bound on the distance between these shapes that can be realized, meaning that the distance between two mass distributions is a closed set along its lower bound. If the bodies are convex this follows trivially.
\subsubsection{Finite Density Sphere Restriction}
This paper mainly focuses on the sphere-restriction of the Full-Body problem, where all of the bodies have finite, constant densities and spherical shapes defined by a diameter $d_i$. This allows for considerable simplification of the mutual potentials, although the rotational kinetic energy, moments of inertia and angular momentum of the systems are still tracked. In this case the moment of inertia of a constant density sphere is $m_i d_i^2 / 10$ about any axis and the minimum distance between two bodies will be $d_{ij} = (d_i + d_j)/2$. The resultant quantities for these systems are
\begin{eqnarray}
T & = & \frac{1}{2M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( \dot{\bfm{r}}_{ij}\cdot\dot{\bfm{r}}_{ij} \right) + \frac{1}{2} \sum_{i=1}^N \frac{m_i d_i^2}{10} {\Omega}^2_i \\
U & = & - {\cal G} \sum_{i=1}^{N-1} \sum_{j=i+1}^N \frac{m_i \ m_j}{|\bfm{r}_{ij}|} \\
\bfm{H} & = & \frac{1}{M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( {\bfm{r}}_{ij}\times\dot{\bfm{r}}_{ij} \right) + \sum_{i=1}^N \frac{m_i d_i^2}{10} \bfm{\Omega}_i
\end{eqnarray}
with the two different versions of the moment of inertia
\begin{eqnarray}
I_p & = & \frac{1}{M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( {\bfm{r}}_{ij}\cdot{\bfm{r}}_{ij} \right) + \frac{3}{2} \sum_{i=1}^N \frac{m_i d_i^2}{10} \\
I_{H} & = & \frac{1}{M} \sum_{i=1}^{N-1}\sum_{j=i+1}^N m_i m_j \left( r_{ij}^2 - (\hat{\bfm{H}}\cdot\bfm{r}_{ij})^2 \right) + \sum_{i=1}^N \frac{m_i d_i^2}{10}
\end{eqnarray}
\subsection{Energy Dissipation Interaction Models}
Implicit in these discussions, although not explicitly incorporated into the interaction models, are the dissipative effects of surface Coulomb friction and the tidal distortion of gravitationally attracting bodies. These physical effects serve dual purpose, in that they will tend to synchronize collections of bodies, either resting on each other or orbiting each other, and will also dissipate excess energy in the system.
No surface interaction models between the finite density bodies is considered other than Coulomb friction between surfaces. This is needed in order for a resting collection of particles to dissipate relative motion between each other when in contact and thus also represents one possible mode of energy dissipation. Inclusion of this notional model ensures that contact configurations will, when reduced to their minimum energy state, all rotate at a common rate. It is possible to explicitly incorporate additional surface potentials between particles in close proximity, such as the Lennard-Jones potential that models van der Waals cohesive forces \cite{castellanos}. These extensions are not considered here, however.
It is also possible to dissipate energy and synchronize spin rates even if bodies are not in contact. Tidal distortions arising from relative motion between gravitationally attracting bodies will also transfer angular momentum across the system and cause the dissipation of energy. Even if these effects are small, as they can be for asteroidal bodies\cite{goldreich}, they are pervasive and will cause continual dissipation of energy for systems that are not in a relative equilibrium state.
The ubiquity and pervasiveness of energy dissipation in the solar system and its role in the long-term evolution of bodies of all sizes motivates the main question concerning the minimum energy states for celestial mechanics systems at a given value of angular momentum.
\section{The Sundman Inequality, Amended Potential, and Minimum Energy Function}
\subsection{The Sundman Inequality}
The fundamental result for this study is found by application of the Sundman Inequality. For generality it is applied to the integral form of the angular momentum vector, given in Eqn.\ \ref{eq:hint}. As this is not the usual form of the equations, a descriptive proof is given that relies on the Cauchy-Schwarz Inequality.
\begin{theorem}
{ $H^2 \le 2I_H T \le 2I_p T$ across all of the differential mass formulations defined in Eqn.\ \ref{eq:dmass} and system moments of inertia defined in Eqns.\ \ref{eq:ipint} and \ref{eq:ih}. }
\end{theorem}
The outermost inequality is the usual Sundman Inequality, but the sharper limits are new and are distinct for the full body problem.
\begin{proof}
First prove the inequality $I_H \le I_p$, which establishes the ordering on the right. First ignore the center of mass terms and only consider the moments of inertia. Taking each term independently leaves $\hat{\bfm{H}}\cdot\bfm{A}_i\cdot\bfm{I}_i\cdot\bfm{A}_i^T\cdot\hat{\bfm{H}} \le \frac{1}{2} \mbox{Trace}(\bfm{I}_i)$. If $I_1 \le I_2 \le I_3$ are the principal moments of inertia of $\bfm{I}_i$, then $\hat{\bfm{H}}\cdot\bfm{A}_i\cdot\bfm{I}_i\cdot\bfm{A}_i^T\cdot\hat{\bfm{H}} \le I_3$. Thus the inequality reduces to $I_3 \le \frac{1}{2}\left(I_1 + I_2 + I_3\right)$ or $I_3 \le I_1 + I_2$. However, this inequality and all of its permutations are a fundamental property of mass distributions and moments of inertia.
Next, consider the center of mass terms, leading to the following inequality for each term: $r_i^2 - (\hat{\bfm{H}}\cdot\bfm{r}_i)^2 \le r_i^2$, which can be trivially shown to be true.
To finish, consider the inequality $H^2 \le 2TI_H$, which is not the usual Sundman Inequality, using the general mass integral form of the angular momentum.
First recall that $\bfm{H} = H \hat{\bfm{H}} = \int_{\cal B} \bfm{r}\times\dot{\bfm{r}} \ dm$. Dotting both sides by the (constant) unit vector aligned with the angular momentum vector yields the equality
\begin{eqnarray}
H & = & \int_{\cal B} \hat{\bfm{H}} \cdot (\bfm{r}\times\dot{\bfm{r}}) \ dm \\
& = & \int_{\cal B} \dot{\bfm{r}}\cdot(\hat{\bfm{H}} \times \bfm{r}) \ dm
\end{eqnarray}
Now apply the triangle inequality to the integral to find:
\begin{eqnarray}
\int_{\cal B} \dot{\bfm{r}}\cdot(\hat{\bfm{H}} \times \bfm{r}) \ dm & \le & \int_{\cal B} \left| \dot{\bfm{r}}\right| \left| \hat{\bfm{H}}\times\bfm{r}\right| dm
\end{eqnarray}
Squaring the original term, equal to $H^2$, and applying the Cauchy-Schwarz inequality yields the main result.
\begin{eqnarray}
H^2 & \le & \left[ \int_{\cal B} \left| \dot{\bfm{r}}\right| \left| \hat{\bfm{H}}\times\bfm{r}\right| dm \right]^2 \\
\left[ \int_{\cal B} \left| \dot{\bfm{r}}\right| \left| \hat{\bfm{H}}\times\bfm{r}\right| dm \right]^2 & \le & \left[ \int_{\cal B} (\hat{\bfm{H}}\times\bfm{r})\cdot(\hat{\bfm{H}}\times\bfm{r}) \ dm \right] \left[ \int_{\cal B} \dot{{r}}^2 \ dm \right]
\end{eqnarray}
But $\int_{\cal B} \dot{{r}}^2 \ dm = 2T$ and $\int_{\cal B} (\hat{\bfm{H}}\times\bfm{r})\cdot(\hat{\bfm{H}}\times\bfm{r}) \ dm = \hat{\bfm{H}}\cdot \int_{\cal B} -\tilde{\bfm{r}}\cdot\tilde{\bfm{r}} \ dm \cdot\hat{\bfm{H}} =\hat{\bfm{H}}\cdot\bfm{I}\cdot\hat{\bfm{H}} = I_H$. Thus, $H^2 \le 2 T I_H$.
\end{proof}
Note that for the point mass density distribution $I_p = I_H$ if all of the bodies lie in a single plane perpendicular to $\hat{\bfm{H}}$, as the moments of inertia are identically zero for a point mass.
\subsection{The Minimum Energy Function and the Amended Potential}
The Sundman Inequality provides an important, and sharp, lower bound on the system energy for a given angular momentum. The derivation of this is simple, but the result has not been extensively used.
\begin{theorem}
The total system energy $E = T+U$ is bounded below by the minimum energy function, defined as ${\cal E} = \frac{H^2}{2I_H} + U$, which is only a function of the system total angular momentum and the relative configuration of the components within in the system. Thus, given a total angular momentum for the system, $H$, the system energy is constrained by ${\cal E} \le E$. Equality of the system energy and the minimum energy function occurs if and only if the minimum energy function is stationary with respect to all possible variations, and corresponds with the system being in a relative equilibrium.
\end{theorem}
\begin{proof}
In the updated Sundman Inequality, $H^2 \le 2TI_H$, replace the kinetic energy with $T = E - U$, where $E$ is the total system energy and $U$ is the gravitational potential of the system. Rearrangement of the terms yields the result $\frac{H^2}{2I_H} + U \le E$, and also defines the minimum energy function. From Eqns.\ \ref{eq:ih} and \ref{eq:udef} note that the minimum energy function is only a function of the system configuration relative to itself, and does not involve any externally defined reference points.
The relative equilibrium result arises from a theorem by Smale \cite{smaleI, smaleII}, proven in a more direct manner by Arnold in \cite{arnoldIII}. First, note that the minimum energy function is related to the ``amended potential'' (with possible differences depending on the form of the polar moment of inertia used) which has found many uses in the derivation and development of constraints on motion in the point mass $N$-body problem. In \cite{smaleII} it is proven that the stationary points of the amended potential correspond to relative equilibria of the corresponding point-mass $N$-body system at a given level of angular momentum. In \cite{simo_marsden} the proof is extended to more general dynamical systems with symmetry, which contains the Full Body system (as established in \cite{cendra_marsden})\footnote{Note that the initial, direct application of the Cauchy-Schwarz inequality yielded a polar moment of inertial $I_p$. If this moment were used to derive the minimum energy function it would no longer equal the amended potential except for the special case of a point-mass system. The modified version of the polar moment, $I_H$ is seen to precisely yield the amended potential when finite density considerations are taken into account.}.
To finish, equality of $E$ and ${\cal E}$ is proven at a relative equilibrium. A collection of bodies in a relative equilibrium will all rotate at a constant rate and maintain a constant polar moment of inertia \cite{saari_constant_polar}. This implies that the angular momentum vector is an eigenvector of the locked inertia matrix and the value $I_H$ is an eigenvalue of this matrix (see also \cite{wang, maciejewski, scheeres_gravgrad}). Thus the total angular momentum is $\bfm{H} = \bfm{I}\cdot\bfm{\Omega} = I_H \Omega\hat{\bfm{H}}$ and the kinetic energy is $T = \frac{1}{2} \bfm{\Omega}\cdot\bfm{I}\cdot\bfm{\Omega} = \frac{1}{2} I_H \Omega^2$. Substituting into the minimum energy function yields ${\cal E} = I^2 \Omega^2 / (2I_H) + U = I \Omega^2 / 2 + U = T + U = E$.
Now consider the case when the system is not in a relative equilibrium yet the minimum energy function equals the system energy, $\frac{H^2}{2I_H} + U = E$. Since the system is not in a relative equilibrium, the minimum energy function (amended potential) is not at a stationary value. Thus, it is possible to take an allowable variation in the configuration to increase its value, or ${\cal E} + \delta{\cal E} > {\cal E} = E$. However, then the Sundman Inequality is violated, forming a contradiction.
\end{proof}
\begin{corollary}
The energy is bounded by an additional function as: $\frac{H^2}{2I_p} + U \le \frac{H^2}{2I_H} + U \le E$.
\end{corollary}
\begin{proof}
This follows immediately from the inequality $I_H \le I_p$.
\end{proof}
\subsection{Existence of Global Energy Extrema}
Given the defined minimum energy function a rigorous approach can be formulated to the original question of what the global minimum energy configurations of a celestial mechanics system are. This, it turns out, is strongly dependent on the density distribution assumed. The following two theorems give explicit results for point mass and finite density distributions.
\begin{theorem}
For the Point Mass Density Distribution: \\
i) the minimum energy function is undefined for $N=1$, \\
ii) the minimum energy function has a unique global minimum for $N=2$, \\
iii) the minimum energy function does not have a global minimum, or even a local minimum, for $N\ge 3$.
\end{theorem}
\begin{proof}
{\it i}) For $N=1$, $I_p = 0$, $U = 0$ and the angular momentum is identically zero for the assumed barycentric coordinate frame. Hence the minimum energy function cannot be constructed. \\
{\it ii)} For $N=2$ a purely constructive proof is given. At $N=2$ the minimum energy function terms are
\begin{eqnarray}
I & = & \frac{m_1 m_2}{m_1+m_2} r^2 \\
U & = & - \frac{{\cal G} m_1 m_2}{r}
\end{eqnarray}
leading to
\begin{eqnarray}
{\cal E} & = & \frac{H^2(m_1+m_2)}{2 m_1 m_2 r^2} - \frac{{\cal G} m_1 m_2}{r}
\end{eqnarray}
and is defined completely by one degree of freedom, the distance between the two mass points. Figure \ref{fig:PM_Ed} shows a generic graph of this function, and it is clear that only a single extremum exists.
\begin{figure}[ht!]
\centering
\includegraphics[scale=0.18]{figures/PM2_Ed}
\includegraphics[scale=0.18]{figures/PM2_EH2}
\caption{A generic graph of the minimum energy function as a function of the distance (left) and of the angular momentum (right). For each angular momentum there is a single extrema, a minimum, which corresponds to a unique circular orbit.}
\label{fig:PM_Ed}
\end{figure}
Taking the variation of ${\cal E}$ only involves this term and yields
\begin{eqnarray}
\delta{\cal E} & = & \frac{H^2(m_1+m_2)}{m_1 m_2 r^3} \left[ - 1 + \frac{{\cal G} (m_1 m_2)^2 r}{H^2(m_1+m_2)}\right] \delta r
\end{eqnarray}
Setting this to zero for all $\delta r$ leads to a unique solution for $r^* \in (0,\infty)$
\begin{eqnarray}
r^* & = & \frac{H^2 (m_1+m_2)}{{\cal G} (m_1 m_2)^2}
\end{eqnarray}
The corresponding energy is
\begin{eqnarray}
{\cal E}^* & = & - \frac{{\cal G}^2 (m_1 m_2)^3}{2H^2(m_1+m_2)}
\end{eqnarray}
and becomes arbitrarily large as $H^2\rightarrow 0$ and goes to 0 as $H^2 \rightarrow \infty$. Figure \ref{fig:PM_Ed} shows a generic plot of this relation, note that at every angular momentum there is only a single energy configuration.
To test for a minimum take the second variation and substitute the nominal solution to find
\begin{eqnarray}
\delta^2{\cal E} & = & \frac{{\cal G} (m_1 m_2)}{r^3}
\end{eqnarray}
which is strictly positve, and hence the relative equilibrium is at least a local minimum of the energy function. To prove that this is a global minimum, it can be shown that ${\cal E} - {\cal E}^* \ge 0$ for all $r$. To establish this note that this inequality defines a quadratic equation in $1/r$ which can be explicitly factored to show that ${\cal E}^*$ is a global minimum at a given value of angular momentum. In terms of usual orbital mechanics this global minimum is equal to a circular orbit.
{\it iii)}
For $N\ge 3$, note that the relative equilibrium are all stationary values of ${\cal E}$, and that ${\cal E} = E$ at the relative equilibrium. By definition these are all central configurations with a specific system rotation rate so that they remain in a relative equilibrium, and have been analyzed extensively by many authors. Moeckel provides a proof that the second variation of the energy at a fixed value of angular momentum is indefinite for all central configurations of the $N\ge3$ problem \cite{moeckel_central}. Thus, there always exist both positive and negative eigenvalues of this second variation, meaning that relative equilibrium are never even local minima of the minimum energy function, which is equal to the energy at a fixed level of angular momentum at a stationary point\footnote{This is not surprising, as if the second variation were definite it would be trivial to constructively prove stability for at least some relative equilibria -- implying that the KAM theorem would not be needed.}. This result removes any and all central configurations and related relative equilibria from consideration as minimum energy configurations.
To completely establish that there are no global minimum it can be shown that the minimum energy function is unbounded from below for any given $H^2 \in (0,\infty)$. Again this can be done by construction. For any $N\ge3$ consider a random spatial distribution of the point masses, but allow two of them, $m_1$ and $m_2$ to be in close proximity, so that their relative distance is $r_{12} = \epsilon$. Then the minimum energy function can be split into two terms, ${\cal E} = {\cal E}' - \frac{{\cal G}m_1 m_2}{\epsilon}$, where ${\cal E}'$ is finite as long as at least one pair of bodies are not arbitrarily close to each other. Then, as the bodies are point masses let $\epsilon \rightarrow 0$ to find explicitly that ${\cal E}\rightarrow -\infty$. Thus the minimum energy function is unbounded from below for an allowable configuration and hence cannot have a global minimum across all possible system configurations.
\end{proof}
For a {\it mechanics} system, this lack of a global minimum energy configuration is problematic, as all mechanical systems are expected to have minimum energy states. For example, this implies that for the three body problem with $H^2 > 0$ there are two final states in the presence of mechanical energy dissipation. First is that one body is ejected, leaving a two-body system and a solo body, with at least one with a well defined minimum energy state. Second is that the system remains bound and, under energy dissipation, two of the components eventually spiral towards each other until a singularity occurs. This latter system is unacceptable in a Newtonian system, as the body interactions will go beyond the validity of Newtonian gravity and physics at some point during their final spiral.
The problem with this picture lies in the point mass density distribution traditionally used for celestial mechanics systems. Now consider minimum energy configurations for the finite density distribution.
\begin{theorem}
The minimum energy function for an $N$-body finite density distribution has a global minimum energy state for every level of angular momentum and for all $N \ge 1$.
\end{theorem}
\begin{proof}
This theorem can be proven by showing that all of the mutual potentials $U_{ij}$ and $H^2/(2I_H)$ have global extrema across their domains. Then the sum of these functions will have global extrema across their combined domains. \\
For $N=1$ mutual gravitational potentials are undefined (as self potentials are neglected) and $U = 0$. \\
For $N\ge 2$ mutual potentials between each of the bodies in a collection are continuous functions of their domain, which is the relative position vector $\bfm{r}_{ij}$ and the relative attitude, as can be specified by the transformation dyadic $\bfm{A}_{ij}$. Let us break up the relative position vector into the unit vector and its magnitude, $\bfm{r}_{ij} = r_{ij} \hat{\bfm{r}}_{ij}$. Then the elements $\hat{\bfm{r}}_{ij}$ and $\bfm{A}_{ij}$ are defined over compact spaces, since these live in SO(2) and SO(3). For small values of $r_{ij}$ there may be additional constraints on the allowable contact locations and orientations between the bodies, but these are all continuously connected to the domain of interest (per the earlier definitions) and define closed sets. For $r_{ij}$, the lower limit of separation was defined earlier to be $d_{ij}$. Thus the domain $r_{ij} \in [d_{ij}, \infty)$, and is only closed at its lower end. Also note that all mutual potential reach a finite limit of 0 for $r_{ij} \rightarrow \infty$. Thus, a one-point compactification can be made to make the interval over which $r_{ij}$ is defined compact. Note that $U_{ij}$ is a well defined and continuous function across this domain. Then by the extreme value theorem $U_{ij}$ has a global minimum and maximum.
Now consider the function $H^2 / (2 I_H)$. The moment of inertia is always $> 0 $ for a finite density distribution, as even when $N=1$ it equals the moment of inertia about the direction $\hat{\bfm{H}}$, always non-zero by definition. As any $r_{ij}\rightarrow\infty$ the function $1/I_H \rightarrow 0$, and thus is well defined. As before, one can argue that $1/I_H$ is continuous over its interval of definition and that its domain is compact (or can be made compact). Thus, by the extreme value theorem $H^2 / (2I_H)$ has a global minimum and maximum.
To complete the proof, trivially note that the addition of continuous functions defined over a common compact domain remains continuous. Thus the extreme value theorem applies to the minimum energy function ${\cal E} = \frac{H^2}{2I_H} + \sum_{i=1}^{N-1}\sum_{j=i+1}^N U_{ij}$, and it has a global minimum and maximum.
\end{proof}
\section{Minimum Energy Configurations in the \\
Spherical Full Body Problem}
Having proven existence for the main result, its implications can be explored using the simplest possible extension of the point-mass celestial mechanics problem to the finite density case, which is the spherical full body problem. See previous works by \cite{wang, maciejewski, scheeres_F2BP, scheeres_gravgrad, fahnestock_icarus, scheeres_F2BP_planar} for more detailed Celestial Mechanics discussions of the Full 2-Body problem for non-spherical bodies. The following analysis strives to be more complete by analyzing all possible options, enabled in part by the simpler model of interaction that the spherical restriction allows.
\subsection{$N=1$}
The mutual potential is identically zero and the moment of inertia of the sphere is equal through any axis through its center. Thus the minimum energy function takes on the simple form
\begin{eqnarray}
{\cal E} & = & \frac{H^2}{m d^2/5}
\end{eqnarray}
and has no degrees of freedom. Thus this system is always in a minimum energy state. Had the sphere restriction not been applied, then the minimum energy state would have the body reoriented so that the maximum moment of inertia would lie along $\hat{\bfm{H}}$ \cite{burns_safronov}.
\subsection{$N=2$}
Now revisit the two-body problem for the case of finite density, focusing on relative equilibrium, local minimum configurations and globally minimum configurations. This problem shows that extensions to finite density mass distributions fundamentally changes the energy structure of this problem and yields the non-intuitive result that there can be multiple circular orbits at a given angular momentum and that there can be unstable circular orbits.
For generality in the following the two bodies have different masses, $m_1$ and $m_2$, and diameters, $d_1$ and $d_2$.
\begin{theorem}
\label{thm:N2orb}
There are 0, 1 or 2 orbital relative equilibria for the 2-body problem as a function of angular momentum. When there are 2 orbital equilibria, one is energetically stable and the other is energetically unstable.
\end{theorem}
\begin{proof}
The potential and polar moments of inertia are:
\begin{eqnarray}
U & = & -\frac{{\cal G}m_1 m_2}{r} \\
I_H & = & \frac{m_1 m_2}{m_1 + m_2} r^2 + I_s \\
I_s & = & \frac{1}{10} \left( m_1 d_1^2 + m_2 d_2^2\right) \\
r & \ge & \frac{d_1 + d_2}{2} = d_{12}
\end{eqnarray}
Normalize this system by
\begin{eqnarray}
\overline{r} & = & r / d_{12} \\
\overline{H} & = & \frac{H\sqrt{ m_1 + m_2}}{\sqrt{{\cal G} d_{12}} m_1 m_2} \\
\overline{I}_s & = & \frac{m_1 + m_2}{ d_{12}^{2} m_1 m_2} I_s \\
\overline{U} & = & \frac{U d_{12}}{{\cal G} m_1 m_2}
\end{eqnarray}
In the following the $\overline{(-)}$ notation will be dropped with the assumption that all quantities are normalized as given above.
With this normalization the range of allowable distances between the particles is defined by $r \ge 1$.
The minimum energy function is then
\begin{eqnarray}
{\cal E} & = & \frac{H^2}{2\left( r^2 + I_s\right)} - \frac{1}{r}
\end{eqnarray}
This is a single degree of freedom function so relative equilibria are found by taking a variation with respect to $r$
\begin{eqnarray}
\delta{\cal E} & = & \left[ - \frac{H^2}{\left( r^2 + I_s\right)^2} + \frac{1}{r^3}\right] r \delta{r}
\end{eqnarray}
Setting the variation equal to zero yields an equation for $r$:
\begin{eqnarray}
r^4 - {H}^2 r^3 + 2 {I}_s r^2 + {I}_s^2 & = & 0 \\
\end{eqnarray}
Applying the Descartes rule of signs the polynomial either has 2 or no solutions, a drastic change from the point-mass problem which always had a single solution\footnote{If ${I}_s \rightarrow 0$ the single solution from the point-mass case is recovered.}. Of particular interest is the point where the transition from zero to 2 solutions occurs, which will be a double root of this polynomial. Set the derivative to zero to find
\begin{eqnarray}
\left( 4 r^2 - 3 {H}^2 r + 4 {I}_s \right) r& = & 0
\end{eqnarray}
Evaluating both equations simultaneously shows that the bifurcation from zero to two solutions occurs at
\begin{eqnarray}
r^* & = & \sqrt{3{I}_s} \label{eq:rstar} \\
{H}^2 & = & \frac{16}{3\sqrt{3}} \sqrt{{I}_s} \label{eq:Hstar}
\end{eqnarray}
Note that the minimum value of ${I}_s$ occurs when $m_1 = m_2$ which leads to $d_1 = d_2$ under an equal density assumption, which leads to ${I}_s \ge 0.4$. Thus, $r^* \ge \sqrt{1.2} > 1$ for all mass values with constant density and the bifurcation structure for $N=2$ is robust and always occurs when the bodies are separate from each other.
An alternative approach is to solve for ${H}^2$ as a function of $r$, yielding
\begin{eqnarray}
{H}^2 & = & \frac{(r^2 + {I}_s)^2}{r^3} \label{eq:N2HofR}
\end{eqnarray}
This has a minimum point, corresponding to the bifurcation of the roots to the equation given above in Eqns.\ \ref{eq:rstar} and \ref{eq:Hstar}.
For higher values of $H$ the system will then have two roots, one to the right of $r^*$ (the outer solution) and the other to the left (the inner solution). At the bifurcation point the system technically only has one solution.
The stability of each of these solutions is determined by inspecting the second variation of ${\cal E}$ with respect to $r$, evaluated at the relative equilibrium. Taking the variation and making the substitution for $H^2$ from Eqn.\ \ref{eq:N2HofR} yields
\begin{eqnarray}
\delta^2{\cal E} & = & \left[ \frac{4}{(r^2+{I}_s)} - \frac{3}{r^2} \right] \frac{(\delta r)^2}{r}
\end{eqnarray}
Checking for when $\delta^2{\cal E} > 0$ yields the condition $r^2 > 3{I}_s$. Thus, the outer solution will always be stable while the inner solution will always be unstable, with the relative equilibria occurring at a minimum and maximum of the minimum energy function, respectively.
\begin{figure}[ht!]
\centering
\includegraphics[scale=0.28]{figures/N2locus}
\caption{Locus of orbital equilibria across a range of angular momentum values. Plotted is the minimum energy function versus the system configuration, $r$, the distance between the two particles. This plot assumes equal masses and sizes of the two particles. For clarity, equilibrium solutions are shown below the physical limit $r \ge 1$.}
\label{fig:N2locus}
\end{figure}
Figure \ref{fig:N2locus} shows characteristic energy curves and the locus of equilibria for different levels of angular momentum for the spherical full 2-body problem. This should be contrasted with
the energy function for the point mass 2-body problem, shown in Fig.\ \ref{fig:PM_Ed}, which only has one relative equilibrium.
\end{proof}
\begin{theorem}
\label{thm:N2rest}
There are 0 or 1 unique resting relative equilibrium configurations as a function of angular momentum. When it exists, the resting equilibrium is energetically stable.
\end{theorem}
\begin{proof}
There is only a single unique contact configuration for the $N=2$ case, with the two bodies touching with their centers of mass a distance $d_{12}$ apart. In normalized coordinates the minimum energy function is ${\cal E} = \frac{{H}^2}{2(r^2 + {I}_s)} - \frac{1}{r}$ where $r=1$ signifies contact between the bodies. For a contact structure to be a resting equilibrium, any variations in its degrees of freedom must conform to the rigid body constraint (i.e., $\delta r \ge 0$ in this case) and must only cause an increase in the minimum energy function, $\delta{\cal E} \ge 0$. If a decrease in energy occurs, the system will trend away from the configuration along an allowable variation in the degree of freedom. Taking the variation in the minimum energy function at the contact configuration of $r=1$ yields
\begin{eqnarray}
\delta{\cal E} & = & \left[ 1 - \frac{{H}^2}{(1 + {I}_s)^2}\right] \delta r
\end{eqnarray}
Since $\delta r \ge 0$ is the only variation allowed, the relative equilibrium will exist when $1 - \frac{{H}^2}{(1 + {I}_s)^2} \ge 0$. Thus if ${H} \le 1 + {I}_s$ the static resting equilibrium will exist and if ${H} > 1 + {I}_s$ there will be no resting equilibrium.
When the resting equilibrium exists, any variation in the degree of freedom will only increase the energy, meaning that its variation is positive definite and this relative equilibrium is energetically stable.
\end{proof}
\begin{theorem}
The complete bifurcation chart of relative equilibria, minimum energy states, and global minimum energy states of the sphere restricted $N=2$ full body problem as a function of angular momentum follows. The sequence is graphically represented in Figure \ref{fig:N2bifurcation}.
\begin{description}
\item
[$0 \le {H}^2 < \frac{16}{3\sqrt{3}} \sqrt{{I}_s}$] The minimum energy configuration, and the only relative equilibrium, is for the two bodies to be resting on each other with $r = d_{12}$.
\item
[$\frac{16}{3\sqrt{3}} \sqrt{{I}_s} \le {H}^2 < \frac{4\sqrt{{I}_s}(1+{I}_s)}{\sqrt{1+{I}_s} + \sqrt{{I}_s}}$] Two orbital relative equilibria bifurcate into existence at a radius $r^* = \sqrt{3{I}_s}$ and grow to larger and smaller values of $r$. The outer solution is at a local minimum of the energy, and thus is stable. The inner solution is at a local maximum of the energy, and thus is unstable. The global minimum energy configuration remains the contact solution.
\item
[$\frac{4\sqrt{{I}_s}(1+{I}_s)}{\sqrt{1+{I}_s} + \sqrt{{I}_s}} \le {H}^2 < (1+{I}_s)^2$] The situation is the same as above, except now the outer orbital solution becomes the global minimum, and the resting solution becomes a local minimum.
\item
[$(1+{I}_s)^2 = {H}^2$] The inner orbital solution collides with the resting solution and they annihilate each other. This can be called the ``fission'' condition for the resting configuration, as it is the point beyond which the resting condition no longer exists and the particles that were resting on each other are forced to enter orbit about each other.
\item
[$(1+{I}_s)^2 < {H}^2$] Beyond the fission limit there is only one relative equilibrium, the outer orbital solution, which is the minimum energy configuration.
\end{description}
\end{theorem}
\begin{figure}[ht!]
\centering
\includegraphics[scale=0.18]{figures/N2bifurcation}
\includegraphics[scale=0.18]{figures/N2diagram}
\caption{The energy-angular momentum diagram of relative equilibria (left) and the transition diagram for increasing and decreasing angular momentum (right) in the Full 2-body problem for equal sized bodies.}
\label{fig:N2bifurcation}
\end{figure}
\begin{proof}
Theorems \ref{thm:N2orb} and \ref{thm:N2rest} outline the values of ${H}^2$ at which the relative equilibria exist and what their stability properties are. The main items to prove in the above are {\it i}) the transitions between the global minimum configuration and {\it ii}) the collision of the unstable orbital equilibrium with the resting equilibrium.
{\it i}) To prove this find the value of ${H}^2$ that gives equality of the energy of the resting and orbital configurations and verify the inequalities.
To understand the global minimum transitions compare the energy of the resting and the stable outer orbital configurations with each other. Equating the orbital and the resting minimum energy functions at the same level of angular momentum, it is possible to solve for the angular momentum at which the two functions are equal:
\begin{eqnarray}
{H}^2 & = & \frac{2(1+{I}_s)(r^2+{I}_s)}{r(r+1)}
\end{eqnarray}
To identify the transition point between global minima substitute for the angular momentum as a function of orbit radius (Eqn.\ \ref{eq:N2HofR}) and simplify the resulting equation to find
\begin{eqnarray}
(r-1)\left( r^2 - 2{I}_s r - {I}_s \right) & = & 0 \label{eq:N2eqcond}
\end{eqnarray}
with an obvious root at $r=1$ (discussed in a moment) and a positive real root at $r_s = {I}_s + \sqrt{{I}_s ({I}_s +1)}$. When $r = r_s$ the global minimum will switch from the resting configuration to the outer orbital configuration as angular momentum increases. Substituting into Eqn.\ \ref{eq:N2HofR} and simplifying yields
\begin{eqnarray}
{H}^2_s & = & \frac{4\sqrt{{I}_s}(1+{I}_s)}{\sqrt{1+{I}_s} + \sqrt{{I}_s}}
\end{eqnarray}
Next check the inequalities $\frac{16}{3\sqrt{3}} \sqrt{{I}_s} < \frac{4\sqrt{{I}_s}(1+{I}_s)}{\sqrt{1+{I}_s} + \sqrt{{I}_s}} < (1+{I}_s)^2$. First note that if ${I}_s$ becomes large the inequalities devolve into $\frac{16}{3\sqrt{3}} \sqrt{{I}_s} < 2 {I}_s < {I}_s^2$ and strictly hold for ${I}_s > 64/27$ for the lower inequality and ${I}_s > 2$ for the upper. Also, taking the minimum value of ${I}_s = 0.4$ and inserting it into the relationships yields $1.9474\ldots < 1.9506\ldots < 1.96$. To verify over the remaining interval it suffices to plot and compare the functions over this compact set.
{\it ii}) Note that the fission of the resting configuration will occur when the inner, unstable relative equilibrium touches the value $r = 1$. From Eqn.\ \ref{eq:N2eqcond} note that the energy of the resting and orbital relative equilibria are equal at this point. Evaluating the angular momentum at this point then yields the value ${H}^2 = (1+{I}_s)^2$.
\end{proof}
\begin{corollary}
The $N=2$ transition diagram as a function of angular momentum is discontinuous and exhibits hysteresis. To track the evolution of a system under changing angular momentum will require dynamical analysis at specific transition points.
\end{corollary}
\begin{proof}
In Figure \ref{fig:N2bifurcation} the energy of the relative equilibria is graphed when they are stable as a function of angular momentum for equal-sized bodies (with qualitatively similar results holding across all values). Under increasing or decreasing angular momentum the configuration will remain in a locally stable relative equilibrium after the global minimum configuration has switched to a different configuration. Thus, when the local relative equilibrium becomes unstable or ceases to exist, there is a finite energy difference between the current system state and its minimum energy state. If angular momentum is increased beyond this point this directly implies a period of dynamical interaction during which energy dissipation may occur, eventually leading to the minimum energy state.
At fission the system has a corresponding energy of $\frac{1}{2}( {I}_s - 1)$ and thus at fission the energy can be positive if ${I}_s > 1$. For constant density spheres this will occur when the ratio between the masses is $< 0.204\ldots$ or $> 4.902\ldots$, assuming constant density \cite{scheeres_F2BP_planar}. In \cite{scheeres_F2BP} it is proven that full body systems can escape if their free energy (equal to the energy as defined in this paper) is positive. Thus, across the range of different mass ratios it is even possible for the resulting fission system to eventually escape prior to sufficient energy dissipation for being trapped into the outer orbital configuration. In \cite{pravec_fission} this same mass ratio is found in a class of natural bodies that are expected to have undergone fission in the past.
\end{proof}
For the Spherical Full 2-body problem, these diagrams are notional as the actual dynamics of these systems consist of purely elliptic motions. Indeed, without explicit energy dissipation the excess energy can never be depleted and the system will not evolve. Also, the rotational motion of the spheres are completely decoupled from orbital motion in the rigid body limit. However, these diagrams present the underlying energy structure for fixed levels of angular momentum. Even the slightest departure from rigidity will enable energy dissipation to occur if perturbed from either of the relative orbital equilibria. If perturbed from the stable equilibrium the dissipation will allow the system to settle into its minimum energy state. If perturbed from the unstable equilibrium it will either evolve towards the contact relative equilibrium, the stable orbital one or may escape before it reaches this limit if the total energy is positive.
\subsection{$N=3$}
This discussion is restricted to bodies having equal sizes and densities. Thus, all particles have a common spherical diameter $d$ and mass $m$. A restriction to the planar problem is also made to simplify the analysis, although non-planar configurations can also be analyzed.
For the equal mass, point-mass celestial mechanics problem, there are only two known relative equilibrium configurations, the Euler and Lagrange solutions. When finite densities are incorporated there are at least 6 orbital configurations and a total of 9 distinct relative equilibria. If non-equal size bodies of similar density are considered, there will be additional distinguishable relative equilibria, but this analysis is left for future work.
Given this restriction the polar moment and potential energy take on simpler forms.
\begin{eqnarray}
I_H & = & \frac{m}{3} \left( r_{12}^2 + r_{23}^2 + r_{31}^2 \right) + \frac{3}{10} m d^2 \\
U & = & - {{\cal G}m^2} \left[ \frac{1}{r_{12}} + \frac{1}{r_{23}} + \frac{1}{r_{31}} \right]
\end{eqnarray}
where $m$ is the common mass of each body and $d$ the common diameter. Then $r_{ij} \ge d$ for all of the relative distances. Now introduce some convenient normalizations, scaling the moment of inertia by $m d^2$ and scaling the potential energy by ${\cal G}m^2 / d$. Then the minimum energy function is
\begin{eqnarray}
\overline{\cal E} & = & \frac{\overline{H}^2}{2 \left[ \left(\overline{r}_{12}^2 + \overline{r}_{23}^2 + \overline{r}_{31}^2\right)/3 + 0.3 \right]}
- \left[ \frac{1}{\overline{r}_{12}} + \frac{1}{\overline{r}_{23}} + \frac{1}{\overline{r}_{31}} \right]
\end{eqnarray}
where
\begin{eqnarray}
\overline{\cal E} & = & \frac{{\cal E} d}{{\cal G}m^2} \\
\overline{H}^2 & = & \frac{H^2}{{\cal G}m^3 d}
\end{eqnarray}
and the constraint from the finite density assumption becomes $\overline{r}_{ij} \ge 1$.
In the following the $\overline{-}$ notation is dropped for $r_{ij}$ and $H$, as it will be assumed that all quantities are normalized.
\begin{theorem}
There are a total of 9 unique (under symmetry transformations) relative equilibria that can exist in the $N=3$ sphere restricted full body problem. These are shown in Figure \ref{fig:N3diagram} and described below.
\begin{enumerate}
\item
There are two static resting relative equilibrium configurations, the Euler Resting and Lagrange Resting configurations. The Lagrange Resting configuration is energetically stable whenever it exists while the Euler Resting configuration transitions from unstable to stable as a function of angular momentum while it exists.
\item
There is one dynamic resting relative equilibrium configurations, the ``V'' configuration, which is always energetically unstable when it exists.
\item
There are 4 mixed orbital and resting relative equilibrium configurations, the inner and outer aligned and transverse mixed configurations. The 2 inner configurations and the outer transverse configurations are always energetically unstable. The outer aligned configuration is always stable.
\item
There are 2 purely orbital relative equilibrium configurations, the Lagrange and Euler configurations. These are always energetically unstable.
\end{enumerate}
\end{theorem}
\begin{figure}[ht!]
\centering
\includegraphics[scale=0.28]{figures/N3diagram}
\caption{Diagram of all relative configurations for the Full 3-body problem. Those boxed can also be stable, and each takes a turn as the global minimum energy configuration for a range of angular momentum.}
\label{fig:N3diagram}
\end{figure}
\begin{proof}
\begin{enumerate}
\item
{\bf Static resting configurations}
First consider the static resting configurations, defined as when all of the bodies are in contact and maintain a fixed shape over a range of angular momentum values. When in contact there is only one degree of relative freedom for the system, defined as the angle between the two outer particles as measured relative to the center of the middle particle and shown in Fig.\ \ref{fig:contact_geometry}.
As defined the angle must always lie in the limit $60^\circ \le \theta \le 300^\circ$.
\begin{figure}[ht!]
\centering
\includegraphics[scale=0.18]{figures/contact_geometry}
\includegraphics[scale=0.18]{figures/mixed_geometry}
\caption{Generic description of the planar contact geometry between three equal sized particles (left) and the mixed configuration geometry (right).}
\label{fig:contact_geometry}
\end{figure}
Given the geometric relationships in Fig.\ \ref{fig:contact_geometry}, the minimum energy function can be written as
\begin{eqnarray}
{\cal E}_S & = & \frac{H^2}{2\left[ \left( 2 + 4\sin^2(\theta/2)\right)/3 + 0.3\right]} - 2 - \frac{1}{2\sin(\theta/2)}
\end{eqnarray}
where the $S$ subscript stands for ``Static.''
The first variation is then
\begin{eqnarray}
\delta_{}{\cal E}_S & = & \sin\theta \left[ - \frac{3 H^2}{\left(2.9+4\sin^2(\theta/2)\right)^2} + \frac{1}{8\sin^3(\theta/2)} \right] \delta\theta \label{eq:rest_cond}
\end{eqnarray}
The second variations will be considered on a case-by-case basis. Since this system has a constraint on the angle $\theta$, both the free variations of $\theta$ and the constrained variations when at the limit must be considered.
\paragraph{Euler Rest Configuration}
If $\theta = 0$ the minimum energy function will be stationary. Define this as the Euler Rest Configuration, which consists of all three particles lying in a single line. The stability of this configuration is evaluated by taking the second variation of the energy function and evaluating it at $\theta = 180^\circ$, yielding
\begin{eqnarray}
\delta^2{\cal E}_{S} & = & - \frac{1}{8} \left[ 1 - \frac{24}{(6.9)^2} H^2 \right] \delta\theta^2
\end{eqnarray}
Recall that the stability condition is that the second variation of the energy be positive definite, yielding an explicit condition for stability as $H^2 > 1.98375$, with lower values of $H^2$ being definitely unstable.
The Euler Rest Configuration energy can be specified as a function of angular momentum:
\begin{eqnarray}
{\cal E}_{SE} & = & \frac{H^2}{2\left(2.3\right)} - \frac{5}{2}
\end{eqnarray}
\paragraph{Lagrange Rest Configuration}
Now consider the constrained stationary point with $\theta = 60^\circ$ ($300^\circ$). Define this as the Lagrange Rest Configuration. Here it suffices to evaluate the first variation at the boundary condition, yielding
\begin{eqnarray}
\left. \delta{\cal E}_{S}\right|_{60^\circ} & = & \frac{\sqrt{3}}{2} \left[ - \frac{3H^2}{(3.9)^2} + 1 \right] \delta\theta
\end{eqnarray}
At the $60^\circ$ constraint $\delta\theta\ge 0$ and the Lagrange Rest Configuration will exist and be stable for $H^2 < 5.07$, but beyond this limit an increase in $\theta$ will lead to a decrease in energy and the relative equilibrium will no longer exist. Note that if the $\theta = 300^\circ$ limit is taken, the sign of the first variation switches but the constraint surface is now $\delta\theta \le 0$, and the same results hold.
The Lagrange Rest Configuration energy can be specified as a function of angular momentum:
\begin{eqnarray}
{\cal E}_{SL} & = & \frac{H^2}{2\left(1.3\right)} - 3
\end{eqnarray}
Comparing the energy of these two rest configurations shows that the Lagrange configuration has lower energy for $H^2 < 2.99$ while the Euler configuration has a lower energy above this level of angular momentum.
\item
{\bf Variable Contact Configurations}
In addition to the static resting configurations it is also possible to have full contact configurations which change as the angular momentum varies. These are not fully static, as they depend on having a specific level of angular momentum, generating centripetal accelerations that balance the gravitational and contact forces. For these configurations, as the level of angular momentum varies the configuration itself shifts, adjusting to the new environment. For the $N=3$ case there is only one such ``variable contact'' configuration when restricted to the plane. This particular configuration is always unstable, yet plays an important role in mediating the stability of the other configurations.
\paragraph{``V'' Rest Configuration}
The variable contact configuration that can exist for this system, called the ``V'' Configuration for obvious reasons, yields the final way for a stationary value of the minimum energy function to exist, with the terms within the parenthesis of Eqn.\ \ref{eq:rest_cond} equaling zero. Instead of solving the resulting quartic equation in $\sin(\theta/2)$ it is simpler to evaluate the angular momentum as a function of the system configuration to find
\begin{eqnarray}
H^2 & = & \frac{\left(2.9 + 4\sin^2(\theta/2)\right)^2}{24\sin^3(\theta/2)}
\end{eqnarray}
The range of angular momenta that correspond to this configuration can be traced out by following the degree of freedom $\theta$ over its range of definition.
Thus the V Rest Configuration will exist for angular momentum values ranging from $H^2 = 1.98375$ at $\theta = 180^\circ$ to $H^2 = 5.07$ at $\theta = 60^\circ$. Note that the angles progress from $\theta = 180^\circ \rightarrow 60^\circ$ as the angular momentum increases, and that the limiting values occur when the Euler Rest Configuration stabilizes and the Lagrange Rest Configuration destabilizes. Note that a symmetric family moves from $\theta = 180^\circ$ to $\theta=300^\circ$ at the same levels of angular momentum.
Taking the second variation and evaluating the sign of $\delta^2{\cal E}_{C}$ along the V configuration for arbitrary variations shows that it is always negative definite over the allowable values of $\theta$ and thus that the V Rest Configuration is always unstable.
\item
{\bf Mixed Configurations}
Now consider mixed configurations where both resting and orbital states can co-exist. For $N=3$ there is only one fundamental topology of this class allowed, two particles rest on each other and the third orbits. Further, from simple symmetry arguments two candidate states for relative equilibrium can be identified, a Transverse Configuration where the line joining the two resting particles is orthogonal to the third particle ($\theta = \pm90^\circ$), and an Aligned Configuration where a single line joins all of the mass centers ($\theta = 0, 180^\circ$). To enable a stability analysis a full configuration description of these systems is introduced which requires two coordinates: the distance from the center of the resting pair to the center of the third particle to be $R$, and the angle between the line $R$ and the line joining the resting pair as $\theta$ (see Fig.\ \ref{fig:contact_geometry}).
The distances between the different components can be worked out as
\begin{eqnarray}
r_{12} & = & 1 \\
r_{23} & = & \sqrt{ R^2 - R \cos\theta + 0.25} \\
r_{31} & = & \sqrt{ R^2 + R \cos\theta + 0.25}
\end{eqnarray}
Thus the minimum energy function takes on the form
\begin{eqnarray}
{\cal E}_M & = & \frac{3H^2}{4\left(1.2 + R^2\right)} - 1 \nonumber \\
& & - \frac{1}{\sqrt{ R^2 - R \cos\theta + 0.25}} - \frac{1}{\sqrt{ R^2 + R \cos\theta + 0.25}}
\end{eqnarray}
where the $M$ stands for ``Mixed.''
Taking the variation with respect to $\theta$ yields
\begin{eqnarray}
\delta_\theta{\cal E}_M & = & \frac{R\sin\theta}{2} \times \\
& & \nonumber \left[ \frac{1}{\left( R^2 - R \cos\theta + 0.25\right)^{3/2}} - \frac{1}{\left( R^2 + R \cos\theta + 0.25\right)^{3/2}}\right] \delta\theta
\label{eq:delta_theta_M}
\end{eqnarray}
As expected, the variation is stationary for the Aligned Configuration, $\theta = 0, 180^\circ$, and for the Transverse Configuration, $\theta = \pm90^\circ$.
The variation in the distance yields
\begin{eqnarray}
\delta_R{\cal E}_M & = & \left[ - \frac{3H^2 R}{2(1.2+R^2)^2} + \frac{2R-\cos\theta}{2\left( R^2 - R\cos\theta + 0.25\right)^{3/2}} \right. \nonumber \\
& & \left. + \frac{2R+\cos\theta}{2\left( R^2 + R\cos\theta + 0.25\right)^{3/2}}\right]
\delta R \label{eq:delta_R_M}
\end{eqnarray}
and is discussed in the following.
\paragraph{Transverse Configurations}
First consider the Transverse Configurations with $\theta = \pm90^\circ$. Evaluating the variation of ${\cal E}_M$ with respect to $R$, setting this to zero, and substituting $\theta = 90^\circ$ allows us to solve for the angular momentum explicitly as a function of the separation distance
\begin{eqnarray}
H_{MT}^2 & = & \frac{4(1.2+R^2)^2}{3\left(R^2 + 0.25\right)^{3/2}}
\end{eqnarray}
This function has a minimum value of angular momentum of $H_{MT}^2 \sim 4.002\ldots$ which occurs at $R = \sqrt{2.6}$. This is an allowable value of separation and thus this bifurcation will indeed occur. For higher values of angular momentum there are two relative equilibria, one with separation less than $\sqrt{2.6}$ and the other with separation larger than this. The inner solution touches the other two particles, forming a Lagrange-like configuration, when $R = \sqrt{3}/2$. Substituting this into the above equation for $H_{MT}^2$ shows that this occurs at a value of $5.07$, which is precisely the value at which the Lagrange Rest Configuration becomes unstable. Recall that this was also the value of angular momentum at which point the V Rest Configuration terminated by reaching $60^\circ$. Thus at this value, which is also equal to the Lagrange Orbit Configuration angular momentum at this distance, the inner Transverse Configuration family of solutions terminates. Conversely, the outer Transverse Configuration persists for all angular momentum values above the bifurcation level.
Now consider the energetic stability of this class of relative equilibria. First note that the cross partials, $\delta^2_{\theta R}{\cal E}_M$ are identically equal to zero for the Transverse Configuration. This can be easily seen by taking partials of Eqn.\ \ref{eq:delta_R_M} with respect to $\theta$ and inserting the nominal value $\theta = \pm 90^\circ$. Next, taking the second partial of Eqn.\ \ref{eq:delta_theta_M} with respect to $\theta$ and evaluating it at the nominal configuration yields
\begin{eqnarray}
\delta^2_{\theta\theta}{\cal E}_{MT} & = & \frac{-3 R^2}{2\left( R^2 + 0.25\right)^{5/2}} (\delta\theta)^2
\end{eqnarray}
and $\delta^2_{\theta\theta}{\cal E}_{MT} < 0$. It is not necessary to check further as this tells us that none of the Transverse Configurations are energetically stable.
The explicit energy of the Transverse Configurations is
\begin{eqnarray}
{\cal E}_{MT} & = & \frac{3H^2}{4\left(1.2 + R^2\right)} - 1 - \frac{2}{\sqrt{ R^2 + 0.25}}
\end{eqnarray}
\paragraph{Aligned Configurations}
Now consider the Aligned Configurations with $\theta = 0,180^\circ$. Again solve for the angular momentum as a function of separation
\begin{eqnarray}
H^2_{MA} & = & \frac{2(1.2+R^2)^2}{3R} \left[ \frac{1}{\left(R-\frac{1}{2}\right)^2} + \frac{1}{\left(R+\frac{1}{2}\right)^2} \right]
\end{eqnarray}
Finding the minimum point of this equation as a function of $R$ yields a cubic equation in $R^2$ without a simple factorization. Root finding shows that it bifurcates into existence at a distance of $R = 2.33696\ldots$ with a value of $H_{MA}^2 = 5.32417\ldots$. Again, there is an inner and an outer solution. The inner solution continues down to a distance of $R = 3/2$, where the two groups touch and form an Euler configuration. The value of the angular momentum at this point equals 6.6125 and equals the value at which the Euler Rest Configuration terminates and the Euler Orbit Configuration is born. The outer solution continues its growth with increasing angular momentum.
Now consider the energetic stability of these solutions. Similar to the Transverse Configurations, the mixed partials of the minimum energy function are identically zero at these relative equilibria.
The second partials of Eqn.\ \ref{eq:delta_theta_M} with respect to $\theta$ yields
\begin{eqnarray}
\delta^2_{\theta\theta}{\cal E}_{AM} & = & \frac{R}{2} \left[ \frac{1}{(R-0.5)^3} - \frac{1}{(R+0.5)^3} \right] (\delta\theta)^2
\end{eqnarray}
which is always positive. The second partial of Eqn.\ \ref{eq:delta_R_M} with respect to $R$ is
\begin{eqnarray}
\delta^2_{RR}{\cal E}_{AM} & = & 2 \left[ \frac{9 H^2}{4(1.2+R^2)^3}\left(R^2-0.4\right) \right. \nonumber \\
& & \left. - \frac{1}{(R-0.5)^3} - \frac{1}{(R+0.5)^3} \right] (\delta R)^2
\end{eqnarray}
The resulting polynomial is of high order and is not analyzed.
Alternately, inspecting the graph of this function shows that it crosses from negative to positive at the bifurcation point, as expected. Thus, initially the outer Aligned Configurations are energetically stable while the inner Aligned Configurations are unstable, and remain so until they terminate at the Euler configuration. To make a final check, evaluate the asymptotic sign of the second energy variation. For $R \gg 1$, $H^2_{MA} \sim 4/3R$. Substituting this into the above and allowing $R\gg 1$ again yields $\delta^2_{RR}{\cal E}_{AM} \sim 1/R^3 \delta R^2$, and thus the outer relative equilibria remain stable from their bifurcation on.
The explicit energy of the Aligned Configurations are
\begin{eqnarray}
{\cal E}_{MA} & = & \frac{3H^2}{4\left(1.2 + R^2\right)} - 1 - \frac{1}{(R-0.5)} - \frac{1}{(R+0.5)}
\end{eqnarray}
A direct comparison between ${\cal E}_{MA}$ and ${\cal E}_{MT}$ at the same levels of angular momentum shows that the Aligned Configurations always have a lower energy than the Transverse Configurations. This is wholly consistent with the energetic stability results found throughout.
\item
{\bf Purely Orbital Configurations}
Finally consider the purely orbital configurations for this case. As this is the sphere restricted problem, the orbital relative equilibria will be the same as exist for the point mass problem.
\paragraph{Euler Solution}
For the Euler solution take the configuration where $r_{12} = r_{23} = R$ and $r_{31} = 2R$, $R \ge 1$, reducing the configuration to one degree of freedom. The minimum energy function then simplifies to
\begin{eqnarray}
{\cal E}_{OE} & = & \frac{3H^2}{2\left(6R^2 + 0.9\right)} - \frac{5}{2R}
\end{eqnarray}
Taking the variation of the minimum energy function with respect to this configuration then yields
\begin{eqnarray}
\delta_R{\cal E}_{OE} & = & - \frac{18 H^2 R}{\left(6R^2+0.9\right)^2} + \frac{5}{2R^2}
\end{eqnarray}
Set this equal to zero and solving for the corresponding angular momentum to find
\begin{eqnarray}
H_{OE}^2 & = & \frac{5}{36} \frac{\left( 6R^2 + 0.9\right)^2}{R^3}
\end{eqnarray}
It can be shown that there are two orbital Euler configurations for $H_{OE}^2 > 8\sqrt{5}/3$ and none for lower values. The non-existence of solutions at a given total angular momentum occurs due to the coupling of the rotational angular momentum of the different bodies. In this case, however, the lower solutions all exist at $R <1$ and thus are not real for this system. In fact, given the constraint $R\ge 1$ there will be a single family of orbital Euler solutions at $H_{OE}^2 \ge 6.6125$ with corresponding radii ranging from $R = 1\rightarrow\infty$ as $H_{OE}^2 = 6.6125\rightarrow\infty$.
The correspond energy of these Euler solutions as a function of $R$ is
\begin{eqnarray}
{\cal E}_{OE} & = & - \frac{5}{24R^3} \left(6R^2 - 0.9\right)
\end{eqnarray}
Our simple derivation of the orbital Euler solutions only considers one-dimensional variations in the distance. However for a complete stability analysis it would be necessary to consider variations of each component in turn, as the instability in the Euler problem occurs transverse to the line of syzygies. However, the instability of the Euler solutions are well known and do not need to be derived here.
\paragraph{Lagrange Solution}
To find the conditions for the Lagrange solution take the configuration to be $r_{12} = r_{23} = r_{31} = R \ge 1$, again reducing the minimum energy function to a single degree of freedom.
\begin{eqnarray}
{\cal E}_{OL} & = & \frac{3H^2}{2\left(3R^2+0.9\right)} - \frac{3}{R}
\end{eqnarray}
The variation now yields the condition
\begin{eqnarray}
3 R \left[ \frac{1}{R^3} - \frac{3H^2}{\left(3R^2+0.9\right)^2}\right] & = & 0
\end{eqnarray}
which can be solved for the angular momentum of the orbital Lagrange solutions as a function of orbit size
\begin{eqnarray}
H_{OL}^2 & = & \frac{\left(3R^2+0.9\right)^2}{3R^3}
\end{eqnarray}
Again, two solutions exist for $H_{OL}^2 > 16/\sqrt{10}$, however the inner solution has radius $R< 1$ and is not allowed by this model. Thus, again for the constraint $R\ge 1$ there is a single family of Lagrange solution orbits that range from $R = 1\rightarrow\infty$ as $H_{OL}^2 = 5.07\rightarrow\infty$.
The corresponding energy of these Lagrange solutions as a function of $R$ is
\begin{eqnarray}
{\cal E}_{OL} & = & - \frac{1}{2R^3} \left(3R^2 - 0.9\right)
\end{eqnarray}
Note again that the Lagrange solutions are always energetically unstable, and in particular that the equal mass Lagrange solutions are also spectrally unstable. As a final point, note that the energy of the Euler solutions is actually less than the energy of the Lagrange solutions when $R^2 < 63/60$, i.e., when $R$ is near unity. For larger values of $R$ the Lagrange solution is always lower energy.
\end{enumerate}
\end{proof}
\begin{corollary}
The complete bifurcation chart of relative equilibria, minimum energy states, and global minimum energy states of the sphere restricted $N=3$ full body problem as a function of angular momentum follows. It is graphically illustrated in Figure \ref{fig:N3bifurcation2}
\begin{description}
\item
[$0 \le H^2 < 1.98375$] The Lagrange Rest Configuration is the only stable relative equilibria and is thus the minimum energy configuration. The Euler Rest Configuration does exist, but is unstable.
\item
[$H^2 = 1.98375$] The V Rest Configuration bifurcates into existence from the Euler Rest Configuration, and remains unstable throughout its life. The Euler Rest Configuration becomes marginally stable.
\item
[$1.98375 < H^2 < 2.99$] The Euler Rest Configuration becomes energetically stable, but the Lagrange Rest Configuration remains the global minimum energy configuration.
\item
[$H^2 = 2.99$] The Euler and Lagrange Rest Configurations have equal energy and are both the global minimum energy configurations.
\item
[$2.99 < H^2 < 5.07$] The Euler Rest Configuration becomes the global minimum energy configuration, while the Lagrange Rest Configuration retains its energetic stability.
\item
[$H^2 = 4.002\ldots$] The Transverse Mixed Configurations bifurcate into existence at a distance of $R=\sqrt{2.6}$.
\item
[$4.002\ldots < H^2 < 5.07$] The Transverse Mixed Configurations continue into an inner and outer family at higher values of angular momentum. Both of these families are energetically unstable throughout their life. The outer family exists for all higher values of angular momentum.
\item
[$H^2 = 5.07$] The V Rest and the Inner Transverse Mixed Configurations terminate at the Lagrange Rest Configuration. The Lagrange Rest Configuration ceases to exist and transitions into the Lagrange Orbital Configuration.
\item
[$5.07 < H^2 < 5.32417\ldots$] The Euler Rest Configuration is the only stable relative equilibrium and is thus the global minimum energy configuration. The Lagrange Orbital Configuration exists for all higher angular momentum and is always unstable.
\item
[$H^2 = 5.32417\ldots$] The Aligned Mixed Configurations bifurcate into existence at a distance of $R = 2.33696\ldots$.
\item
[$5.32417\ldots < H^2 < 5.65907\ldots$] The Inner and Outer Aligned Mixed Configurations exist. The inner configuration is always energetically unstable throughout its life. The outer configuration is energetically stable and exists for all higher values of angular momentum. The Euler Rest Configuration remains the global minimum energy configuration.
\item
[$H^2 = 5.65907\ldots$] The Euler Rest and Outer Aligned Mixed Configurations have equal energy and are both the global minimum energy configurations.
\item
[$5.65907\ldots < H^2 < 6.6125$] The Outer Aligned Mixed Configuration becomes the global minimum energy configuration, while the Euler Rest Configuration retains its energetic stability.
\item
[$H^2 = 6.6125$] The Inner Aligned Mixed Configurations terminates at the Euler Rest Configuration. The Euler Rest Configuration ceases to exist and transitions into the Euler Orbital Configuration.
\item
[$6.6125 < H^2$] The Outer Aligned Mixed Configuration is the only stable relative equilibrium and is thus the global minimum energy configuration, and remains so for all higher angular momentum. The Euler Orbital Configuration exists for all higher angular momentum and is always unstable.
\end{description}
\end{corollary}
\begin{figure}[ht!]
\centering
\includegraphics[scale=0.28]{figures/N3bifurcation2}
\caption{Graphic depiction of all relative equilibria and their interactions. Green indicates stability while red indicates instability. The global minimum is not denoted in this diagram.}
\label{fig:N3bifurcation2}
\end{figure}
\begin{proof}
The transitions can all be evaluated by comparison between the existence and stability transitions outlined in the previous proof and by direct comparison of energy levels.
\end{proof}
\begin{corollary}
The $N=3$ global and local minimum energy transition diagram as a function of angular momentum is discontinuous and exhibits hysteresis.
Figure \ref{fig:N3bifurcation} tracks the energetically stable configurations as a function of angular momentum, shows where explicit transitions occur, and indicates the excess energy at each transition. To track the evolution of a system under changing angular momentum will require dynamical analysis at specific transition points.
\end{corollary}
\begin{figure}[ht!]
\centering
\includegraphics[scale=0.28]{figures/N3bifurcation}
\caption{Bifurcation diagram showing the energy-angular momentum curves of all stable relative equilibria and their transition paths for increasing and decreasing angular momentum.}
\label{fig:N3bifurcation}
\end{figure}
\begin{proof}
The plot is created by comparing the energies of the different energetically stable configurations and computing energy differences at transitions. Since excess energies exist at different transition points, it will be necessary for the dynamical evolution of the system to be followed to determine actual outcomes.
\end{proof}
\subsection{$N\gg1$}
The paper ends with a few observations for when $N$ is a large number. Simulations of large numbers of grains that gravitationally attract each other and can rest on each other have been explored over the last decade using a variety of simulation methods. The original work in this area was performed by Richardson and collaborators (c.f.\ \cite{richardson_equilibrium}) and used hard-sphere models for surface interactions. More recent models using soft-sphere models for grain interactions have been explored by S\`anchez and Scheeres \cite{sanchez_ApJ}, and in some cases have advantages in terms of tracking the system-wide energy, angular momentum, and internal forces of these systems when there are significant components that rest on each other. Investigations of relative equilibria of the minimum energy function for such large particle simulations could be investigated, but have not been as of yet.
There are several direct observations that can be made for large $N$ systems. First, for non-rotating systems it is expected that the global minimum configuration will tend towards a spherical collection of particles resting on each other. Depending on the presence and strength of friction between the grains, it is also possible that a variety of non-spherical shapes may be stable relative equilibria, thus generalizing the ideal results from continuum mechanics and theory. Likewise, as angular momentum is increased the classical Maclaurin and Jacobi ellipsoids should always be relative equilibria, however one must again note the probable existence of multiple other non-ellipsoidal configurations that are also relative equilibria. This is to be expected, given the presence of such multiple configurations at a given level of angular momentum for particle systems with $N > 2$. Of specific interest are the evolutionary pathways and transitions for these systems as the total angular momentum is increased (or decreased) mimicking the natural evolution of rubble pile asteroids under the YORP effect. Analytical studies of this have been made by Holsapple \cite{holsapple_yorp} and recreated using simulations by S\`anchez \cite{sanchez_icarus}. Of particular interest for these studies are the pathways to fission of large collections of particles. A systematic study of these transitions will be of real interest in understanding the overall evolution and creation of binary bodies among small asteroids, at the least\cite{jacobson_icarus}.
To close this section, a final hypothesis is given.
\begin{hypothesis}
For all $N\ge 2$, energetically stable relative equilibrium configurations of Full Body problems will result in one of two general states:
\begin{enumerate}
\item
All the particles rest on each other in one collection and spin at a uniform rate;
\item
All the particles separate into two collections in a mutually circular orbit about each other with doubly-synchronous rotation.
\end{enumerate}
Further, that all configurations where the particles condensate into 3 or more collections will be energetically unstable and will either eject the excess collections or, under energy dissipation, will condense into one or two collections.
\end{hypothesis}
One possible avenue to the proof of this hypothesis is through the proof for the point mass $N$ body problem that all orbital relative equilibria with $N\ge 3$ are energetically unstable. As for most collections with $N\ge3$ in relative equilibria, the leading order of their gravitational attractions will always represent the point mass $N$ body problem, with higher-order gravitational perturbations arising from the non-spherical configuration of the condensed grains in each collection.
\section{Conclusions}
This paper considers the basic and fundamental question: What are the minimum energy and stable configurations of the $N$-body problem in celestial mechanics for a given value of angular momentum. It is shown that this problem is ill-posed for the classical point mass model of celestial mechanics, but is well posed if all density distributions for these systems are made finite. This stipulation requires that rotational energy and angular momentum also be considered in searching for these stable configurations, and indeed fundamentally changes the problem. Specifically, under this reasonable stipulation the minimum energy solutions of the finite density 2-body problem change fundamentally as compared to the point mass 2-body problem. This question is also rigorously explored for $N=3$ along with a number of hypotheses for the $N\gg1$ problem. In the course of these discussions a number of fundamental questions are identified on the evolution of celestial mechanics systems under an increase of angular momentum, as has been documented to occur in nature.
\section*{Acknowledgements}
Fruitful and motivating conversations with James Montaldi and Manuele Santoprete at a Mathematisches Forschungsinstitut Oberwolfach workshop in August 2011 are acknowledged.
The author acknowledges support from NASA grants NNX11AP24G and NNX10AG53G from the Planetary Geology and Geophysics and the Near Earth Objects Observation programs, respectively.
\newpage \bibliographystyle{plain}
|
\section{Introduction}
Going back to Plato, classification or categorization is the epistemological process that groups objects based on similar properties \cite{plato}. Having primarily biological examples in mind, Aristotle defined categories as discrete entities characterized by properties shared by their members \cite{arist}. Locke in 1690 distinguished between the {\it nominal} and the {\it real essence} of an object \cite{locke}. The nominal essence comes from experience and represents the object's appearance, the real essence represents the object's deeper, constituting features. For instance, the real essence of a material thing is its atomic constitution, because this is the causal basis of all the thing's observable properties \cite{wiki}. A scientific classification is particularly useful if it reflects the {\it real} essence of the objects in question by identifying their underlying common features, from which the more obvious and easily observable {\it nominal} properties follow. Having in mind Locke's concept of real essence, we argue below for a new definition of the class of simple liquids.
Physicists love simple systems. This reflects the fundamental paradigm that in order to capture a given phenomenon, simpler is better. Most classifications in physics are clear-cut, for example the classification of elementary particles into baryons and leptons, whereas classifications in other sciences usually have a wealth of borderline cases. Due to the diversity of molecules it is reasonable to expect a definition of ``simple liquids'' to be of the latter type.
The concept of a simple liquid is old, but it remains central as evidenced by the 2003 book title ``Basic concepts for simple and complex liquids'' \cite{barrat} or the review of non-simple liquids entitled ``Theory of complicated liquids'' from 2007 \cite{kir07}. Generations of liquid-state theorists were introduced to the topic by studying Hansen and McDonald's textbook ``Theory of simple liquids'' \cite{han05}. This book first appeared in 1976 following a period of spectacular progress in the theory of liquids, catalyzed by some of the very first scientific computer simulations.
In Ref. \onlinecite{han05} a simple liquid is defined as a classical system of approximately spherical, nonpolar molecules interacting via pair potentials. This and closely related definitions of liquid simplicity have been standard for many years \cite{fis64,ric65,tem68,ail80,gubbins}. In this definition simple liquids have much in common with the chemists' ``nonassociated liquids'' \cite{chandler}, but there are some differences. Chemists generally regard a liquid as simple even if it consists of elongated molecules, as long as these are without internal degrees of freedom and interact primarily via van der Waals forces. Many physicists would probably disagree. Thus it is far from trivial to ask: What characterizes a ``simple liquid''? More accurately: Given a classical system of rigid bodies with potential energy as a function of the bodies' centers-of-masses and their spatial orientations, is it possible to give a quantitative criterion for how simple the system is? If yes, is simplicity encoded uniquely in the potential energy function or may the degree of simplicity vary throughout the phase diagram?
Recent works identified and described the properties of ``strongly correlating liquids'' \cite{ped08,ped10,effectivetemperature,baileyOrderParameters,paperI,paperII,paperIII,paperIV,paperV,paper00,sch09,ped11}. In these liquids, by definition, the virial $W$ and the potential energy $U$ correlate strongly in their constant-volume thermal-equilibrium fluctuations. Recall that the average virial $\langle W\rangle$ gives the contribution to pressure from intermolecular interactions, added to the ideal-gas term $Nk_BT$ deriving from momentum transport via particle motion (below $p$ is the pressure, $V$ the volume, $N$ the number of particles, $k_B$ Boltzmann's constant, and $T$ the temperature):
\begin{equation}
pV
\,=\,Nk_BT+\langle W\rangle\,.
\end{equation}
The term ``strongly correlating liquid'' refers to the case when the $WU$ correlation coefficient in the $NVT$ ensemble is larger than $0.9$ \cite{paperI}. If angular brackets denote an $NVT$ ensemble average, the correlation coefficient $R$ is defined by
\begin{equation}\label{Rdef}
R
\,=\,\frac{\langle \Delta W \Delta U \rangle}{\sqrt{\langle (\Delta W)^{2} \rangle \langle (\Delta U)^{2} \rangle}}\,.
\end{equation}
An example of a strongly correlating liquid is the standard Lennard-Jones liquid at typical condensed-phase state points, i.e., not far from the solid-liquid coexistence line. Many other systems, including some molecular models, have been shown to be strongly correlating; we refer the reader to papers that derive and document the several simple properties of strongly correlating liquids
\cite{ped08,ped10,effectivetemperature,baileyOrderParameters,paperI,paperII,paperIII,paperIV,paperV,paper00,sch09,ped11}, reviewed briefly in Ref. \onlinecite{ped11}. These properties are summarized in Sec. \ref{prop} after the presentation of the simulation results.
The present work is motivated by developments initiated by recent findings by Berthier and Tarjus \cite{wcabertier1,wcabertier2}. These authors showed that for the viscous Kob-Andersen binary Lennard-Jones mixture \cite{ka1,ka2} the dynamics is not reproduced properly by cutting the potentials at their minima according to the well-known Weeks-Chandler-Andersen (WCA) recipe \cite{wca}. The role of the cutoff was subsequently studied in two papers, showing that placing a shifted-forces cutoff at the first minimum of the pair correlation function gives good results for Lennard-Jones type systems \cite{FCS1,FCS2}. This applies not only at moderate densities, but also at very high densities. Applying the same cutoff to water does not work very well \cite{pal06}. Water is an example of a non-strongly correlating liquid with $R\approx 0$ at ambient conditions, a consequence of water's density maximum \cite{paperI}. These findings led us to speculate whether it is a general property of strongly correlating liquids that the intermolecular interactions may be ignored beyond the FCS without compromising accuracy to any significant extent. The main part of the present paper shows that, indeed, using such an ``FCS cutoff'' gives accurate simulation results if and only if the liquid is strongly correlating.
The paper presents results obtained from computer simulations of 15 different systems, some of which are strongly correlating. We investigate the role of the FCS in determining liquid structure and dynamics. Structure is probed by the radial distribution function (RDF), dynamics by the incoherent or coherent intermediate scattering function (ISF) at the wavevector of the static structure factor maximum. The numerical evidence is clear. By varying the cutoff of the intermolecular forces, we find that in order to get accurate simulation results it is enough to take into account the interactions within the FCS {\it if and only if} the liquid is strongly correlating. In other words, for strongly correlating liquids interactions beyond the FCS are unimportant, and this applies {\it only} for these liquids. At present there is no rigorous argument for this empirical ``FCS property'', but we show that it is consistent with known properties of strongly correlating liquids.
The FCS property of strongly correlating liquids documented below emphasizes further that these are simpler than liquids in general. A number of other simple properties of strongly correlating liquids were reported previously \cite{ped08,ped10,effectivetemperature,baileyOrderParameters,paperI,paperII,paperIII,paperIV,paperV,sch09,ped11}. Altogether, these facts motivate our new definition of liquid simplicity.
Section \ref{sim_sec} presents the results from molecular dynamics simulations and Sec. \ref{sum} summarizes the results. Section \ref{SCL_sec} gives an overview of the many simple properties of strongly correlating liquids, motivating our suggestion that a liquid is to be defined as simple whenever it is strongly correlating at the state point in question. Section \ref{disc} gives a few concluding remarks.
\section{Molecular dynamics simulations of atomic and molecular liquids}\label{sim_sec}
In a computer simulation the intermolecular interactions, which usually extend in principle to infinity, are truncated at some cutoff distance $r_{c}$ beyond which they are ignored \cite{FCS1}. To avoid a discontinuity in the force, which can severely affect the simulation results \cite{FCS1,tildesley}, the simulations reported below use potentials truncated such that the force goes continuously to zero at $r_{c}$. This is done by applying a so-called shifted forces (SF) cutoff \cite{tildesley,kan85,hal08} where, if the pair potential is $v(r)$ and the pair force is $f(r)=-v'(r)$, the shifted force is given by
\begin{equation}
f_{\rm SF}(r)\,=\,
\begin{cases} f(r)-f(r_c) & \text{if}\,\, r<r_c\,, \\
0 &\text{if}\,\, r>r_c\,.
\end{cases}
\end{equation}
This corresponds to using the following pair potential below $r_c$: $v_{\rm SF}(r) = v(r) - v'(r_c) ( r-r_c) - v(r_c)$. Using an SF cutoff gives more accurate results and better numerical stability than using the standard shifted-potential (SP) cutoff \cite{FCS1}. This is so despite the fact that an SF cutoff does not have the correct pair force for any $r$, whereas the pair force is correct below $r_c$ for an SP cutoff. Apparently, avoiding discontinuity of the force at $r_c$ is more important than maintaining the correct force. It was recently discussed why adding a linear term to the pair potential affects neither structure nor dynamics to any significant extent \cite{paperII}. The reason is that, when one nearest-neighbor distance decreases, others increase in such a way that their sum is virtually constant. This argument is exact in one dimension and holds to a good approximation in 3D constant-volume ensemble simulations \cite{paperII} (in constant-pressure ensemble simulations the volume fluctuates and the argument no longer applies). Coulomb interactions have also been treated by the SF cutoff procedure. Although the Coulomb interaction is long-ranged and conditionally convergent, when $r_c$ is sufficiently large an SF cutoff gives results close to those of the standard, much more involved Ewald summation method \cite{fennell,jesper}.
All simulations were performed in the $NVT$ ensemble with periodic boundary conditions and the Nose-Hoover algorithm \cite{nose,hoover,frenkel}. We used the RUMD molecular dynamics package developed in-house, which is optimized for simulations of small systems on state-of-the-art GPU hardware (a few thousand entities) \cite{rumd}. For the molecular models bond lengths were held fixed using the time-symmetrical central difference algorithm \cite{nvttoxvaerd,toxconstraintnph,toxconstraintnve}.
In the following we investigate for several systems whether it is possible to choose an FCS cutoff, i.e., a cutoff at the first minimum of the pair correlation function, and still get the correct physics. We start by studying strongly correlating atomic liquids. Then data are presented for atomic liquids that are not strongly correlating. Finally data are given for two strongly correlating molecular liquids and a water model. Details of the models studied, number of particles, etc, are given in Appendix \ref{pairP}.
\subsection{Three inverse-power-law (IPL) fluids}
We consider first systems with 100\% correlation between virial and potential energy equilibrium fluctuations in the $NVT$ ensemble. It follows from the definition of the virial $W=-1/3\sum{\bf r}_i\cdot\nabla_i U$ \cite{tildesley} that a necessary and sufficient condition for $W$ to correlate perfectly with $U$ is that the potential energy is a so-called Euler homogeneous function of the particle coordinates ${\bf r}_i$. This is the case for systems with inverse power-law (IPL) pair potentials ($v(r)\propto r^{-n}$), but note that potentials with non-trivial angular dependence may also be Euler homogeneous.
We simulated single-component IPL pair-potential systems with exponents $n = 18, 6, 4$ at density $\rho = 0.85$. Each system was simulated at two temperatures. The simulated systems range from very harsh repulsive ($n=18$) to quite soft and long ranged ($n=4$). Each system was simulated with four SF cutoffs. The role of the cutoff is investigated by choosing three different cutoffs: one placed at the first minimum of the RDF, one corresponding to the half height of the RDF from its first maximum to the first minimum, and one placed to the right of the RDF first minimum displaced the same amount as the difference between the first and the second cutoff.
The RDFs $g(r)$ are shown for $n = 18, 6, 4$ in Fig. \ref{IPLgrs}; $n=12$ gives similar results (not shown). The simulations with an SF cutoff at the first minimum of the RDF -- referred to as FCS-cutoff simulations -- give a faithful representation of the structure. The insets show the deviations in RDF as functions of the cutoff between results for an SF cutoff and ``true'' results, quantified by integrating the numerical difference in the pair correlation function. Clearly, deviations increase sharply when the cutoff enters the FCS (blue crosses).
We simulated also the $n=3$ and $n=1$ IPL fluids (results not shown), the latter is usually termed the one-component plasma. For both systems an FCS cutoff does not lead to the correct physics. This is most likely because the potentials are too long ranged for such a short cutoff to make good sense. In fact, both are so long ranged that they do not have a proper thermodynamic limit for which the exponent must be larger than the dimension \cite{fis66}. An indication that an FCS cutoff works poorly when the IPL exponent approaches the dimension is seen for the $n=4$ simulation, for which the $WU$ correlation coefficient for the FCS system starts to deviate significantly from unity. Moreover, almost invisible in the figure is the fact that the $n=4$ pair correlation function's first maximum deviates slightly between FCS and true simulations.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/IPL18_rho0850_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/IPL06_rho0850_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/IPL04_rho0850_gr}
\vspace{-20pt}
\caption{RDFs for single-component {IPL} fluids with exponents $n=18,6,4$, each simulated at two temperatures at density $\rho = 0.85$. The black and orange curves show reference simulation results with large cutoffs representing the ``true'' IPL behavior, the red and green dots give results from simulations with an FCS cutoff (marked by the vertical red dashed lines). The insets quantify the deviations in the RDF from the reference RDF as functions of the cutoff; deviations increase dramatically when the cutoff enters the FCS (blue crosses). In each subfigure the virial/potential-energy correlation coefficient $R$ is given for the simulated cutoff (this quantity is exactly unity for IPL systems with infinite cutoff).}
\label{IPLgrs}
\end{figure}
Figure \ref{IPLFsAs} shows the incoherent intermediate scattering functions (ISFs) for the low-temperature state points of each of the three IPL systems evaluated at the wavevector corresponding to the first maximum of the static structure factor. A good representation of the dynamics is obtained for all systems when the FCS cutoff is used.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/IPL18_rho0850_FsA}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/IPL06_rho0850_FsA}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/IPL04_rho0850_FsA}
\vspace{-20pt}
\caption{Incoherent intermediate scattering functions (ISFs) for the {IPL} fluids at the lowest-temperature state points of Fig. \ref{IPLgrs}.
The black curves give results for a large cutoff, the red crosses for an FCS cutoff (marked by the vertical red dashed lines in Fig. \ref{IPLgrs}).
(a) $n=18$, $T = 0.30$;
(b) $n=6$, $T = 0.15$;
(c) $n=4$, $T = 0.10$.
}
\label{IPLFsAs}
\end{figure}
\subsection{Lennard-Jones type liquids}
Next, we consider the most studied potential in the history of computer simulations, the Lennard-Jones (LJ) pair potential,
\begin{equation}\label{LJ}
v_{\rm LJ}(r) = 4\epsilon \Big[ \Big( \frac{\sigma}{r}\Big)^{12} - \Big(\frac{\sigma}{r}\Big)^{6} \Big]\,.
\end{equation}
Here $\sigma$ and $\epsilon$, define, respectively, the length and energy scale of the interaction (dimensionless units defined by $\sigma=\epsilon=1$ are used below). This potential does not have 100\% virial/potential-energy correlation, but has still quite strong correlations with correlation coefficients $R>0.9$ in the condensed-fluid part of the phase diagram (also in the crystalline phase \cite{paperII}). We studied the single-component LJ (SCLJ) liquid, two generalized 80/20 Kob-Andersen binary LJ (KABLJ) mixtures with repulsive exponent $12$ and attractive exponents $n=4,10$, and the Wahnstrom 50/50 binary LJ (WABLJ) mixture (Fig. \ref{pairKABLJ} and Appendix \ref{pairP} gives model details). The influence of an SF cutoff on simulation accuracy was investigated recently for the SCLJ liquid and the standard KABLJ mixture ($n=6$) \cite{FCS1,FCS2}, but for completeness we include results for the SCLJ system here as well.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/KABLJ_pair_potential}
\caption{The AA-particle generalized Kob-Andersen (KABLJ) pair potentials with fixed repulsive exponent $12$ and three different attractive exponents $n=4,6,10$. The model parameters are: $\epsilon_{AA}=1.00$, $\epsilon_{AB}=1.50$, and $\epsilon_{BB}=0.50$, $\sigma_{AA}=1.12$, $\sigma_{AB}=0.90$, $\sigma_{BB}=0.99$, $m_A=m_B=1$.}
\label{pairKABLJ}
\end{figure}
The role of the cutoff is again investigated by choosing three different cutoffs: one placed at the first minimum of the RDF (red color in Figs. \ref{SCLJgr}-\ref{WaF}), one corresponding to the half height of the RDF from maximum to the first minumum (blue color in Figs. \ref{SCLJgr}-\ref{WaF}), and one displaced to the right of the minimum the same amount as the difference between the first and the second cutoff (green color in Figs. \ref{SCLJgr}-\ref{WaF}).
In Fig. \ref{SCLJgr} RDFs are shown for the {SCLJ} liquid at three different state points. The red colored circles show results from simulations with an FCS cutoff (marked by the vertical red dashed line), the black curves show the corresponding simulations with a large cutoff (reference system). The insets quantify the deviations in the simulated RDF from the reference RDF as a function of the cutoff. The reference RDF of Figs. \ref{SCLJgr}(a) and (b) is clearly represented well using an FCS cutoff, while choosing the cutoff inside the FCS results in significant deterioration. At low densities [Figs. \ref{SCLJgr}(c)], deviations occur between FCS cutoff simulations and the reference system. As mentioned, the {SCLJ} liquid is strongly correlating in large parts of its phase diagram, but as density is lowered, the correlations decrease gradually, and the liquid is no longer strongly correlating at state point (c) where $R = 0.50$. These simulations suggest that only when a liquid is strongly correlating is it possible to ignore interactions beyond the FCS.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCLJ_rho0850_T0700_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCLJ_rho0850_T1000_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCLJ_rho0550_gr}
\caption{RDFs for the single-component Lennard-Jones (SCLJ) liquid at three different state points:
(a) $\rho=0.85$, $T=0.70$ ($R = 0.96$);
(b) $\rho=0.85$, $T=1.00$ ($R = 0.97$);
(c) $\rho=0.55$, $T=1.13$ ($R = 0.50$).
The black curves show reference simulations with large cutoffs, the red dots/curve show results from simulations with an FCS cutoff (marked by the vertical red dashed lines). The insets quantify the deviation in RDF from the reference RDF as functions of the cutoff. At all three state points deviations increase significantly when the cutoff enters the FCS (blue crosses in the insets). For state points (a) and (b), which are strongly correlating ($R>0.9$), an FCS cutoff leads to accurate results. This is not the case for state point (c) that is not strongly correlating.}
\label{SCLJgr}
\end{figure}
Next, we investigated the SCLJ dynamics at the three state points of Fig. \ref{SCLJgr}. The dynamics is studied via the ISF. The ISFs are shown in Fig. \ref{SCLJdyn}; at all state points the dynamics is represented well using an FCS cutoff.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCLJ_rho0850_T0700_FsA}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCLJ_rho0850_T1000_FsA}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCLJ_rho0550_FsA}
\caption{{ISF}s for the {SCLJ} liquid at the state points of Fig. \ref{SCLJgr}. The black curves give the reference cutoff results, the red curves the FCS cutoff results, the blue curves results for a cutoff at the half-height towards the first maximum of the RDF, the green curves results for a cutoff to the right of the minimum.
(a) $\rho=0.85$, $T=0.70$ ($R = 0.96$);
(b) $\rho=0.85$, $T=1.00$ ($R = 0.97$);
(c) $\rho=0.55$, $T=1.13$ ($R = 0.50$).}
\label{SCLJdyn}
\end{figure}
\begin{table}
\begin{tabular}
{|p{1.3cm}||p{0.9cm}|p{0.9cm}|p{0.9cm}||p{2.0cm}|p{2.0cm}|}
\hline
\hline
System & $\rho$ & $T$ & $R$ & $|\Delta RDF|_{max}$ & $|\Delta ISF|_{max}$ \\
\hline
\hline
SCLJ & 0.85 & 1.00 & 0.97 & 1.31$\cdot10^{-2}$ & 5.10$\cdot10^{-3}$ \\
SCLJ & 0.85 & 0.70 & 0.96 & 1.68$\cdot10^{-2}$ & 8.28$\cdot10^{-3}$ \\
SCLJ & 0.85 & 0.65 & 0.96 & 1.63$\cdot10^{-2}$ & 8.96$\cdot10^{-3}$ \\
SCLJ & 0.50 & 1.50 & 0.69 & 11.2$\cdot10^{-2}$ & 7.94$\cdot10^{-3}$ \\
SCLJ & 0.55 & 1.13 & 0.50 & 15.2$\cdot10^{-2}$ & 12.0$\cdot10^{-3}$ \\
\hline
\hline
\end{tabular}
\caption{Simulation results for five state points of the SCLJ liquid. For each state point is given density, temperature, correlation coefficient, maximum deviation from the true RDF using an FCS cutoff, and maximum deviation from the true incoherent structure factor using an FCS cutoff.}
\label{table}
\end{table}
Table \ref{table} summarizes results for five state points of the SCLJ liquid. The third state point is slightly below the triple point, i.e., in a slightly metastable state. The deviations clearly increase as the $WU$ correlations decrease.
We proceed to investigate mixtures of two different particles (A and B) interacting with LJ type potentials. The cutoff used for all three interactions (AA, AB, BB) are placed at the same distance, referring to $\sigma_{AA}$. In Fig. \ref{KABLJgr} reference and FCS cutoff results are shown for the $AA$-particle RDFs of generalized KABLJ mixtures with repulsive exponent $12$ and attractive exponents $n=4,10$. The $n=4$ case is below our definition of a strongly correlating liquid $R>0.9$ (at the simulated state point this system has negative virial), but is still pretty well correlating. For all investigated state points an FCS cutoff gives accurate results. We found the same using the standard repulsive exponent $n=6$ \cite{FCS2} (results not shown).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/KABLJ04_rho1200_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/KABLJ10_rho1200_gr}
\caption{RDFs for generalized KABLJ mixtures with repulsive exponent $12$ and attractive exponents $n=4,10$. The black curves give the reference cutoff results, the red curves the FCS cutoff results. The insets quantify the deviation in RDF from the reference RDF as functions of the cutoff.
(a) $n=4$, $\rho=1.20$, $T=0.40$ ($R = 0.84$);
(b) $n=10$, $\rho=1.20$, $T=1.00$ ($R = 0.97$). }
\label{KABLJgr}
\end{figure}
The A-particle {ISF}s for the state points of Fig. \ref{KABLJgr} are shown in Fig. \ref{KABLJFsA}. For the KABLJ mixture, placing the cutoff inside the FCS (blue curves) fails to reproduce the dynamics properly, whereas it is well approximated using an FCS cutoff (red). Slight deviations are noted for the red curves, an issue considered in Appendix \ref{FCS_sec} that discusses alternatives for delimiting the FCS. Similar results are found for the B particles (results not shown).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/KABLJ04_rho1200_FsA}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/KABLJ10_rho1200_FsA}
\caption{{ISF}s for generalized KABLJ mixtures with repulsive exponent $12$ and attractive exponents $n=4,10$.
The red and black curves give, respectively, results for FCS cutoffs and large reference cutoffs.
(a) $n=4$, $\rho=1.20$, $T=0.40$ ($R = 0.84$);
(b) $n=10$, $\rho=1.20$, $T=1.00$ ($R = 0.97$).}
\label{KABLJFsA}
\end{figure}
We also simulated the Wahnstr{\"o}m 50-50 binary LJ mixture \cite{Wahnstromblj}, finding again that whenever $R>0.9$ structure and dynamics are well reproduced using an FCS cutoff. We do not show these results, but show instead results for the coherent scattering function at one state point (Fig. \ref{WaF}). Again, the FCS cutoff (red crosses) gives the correct dynamics whereas reducing the cutoff further does not give proper results (blue crosses).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/Wahnstrom_rho1296_F}
\caption{Intermediate coherent scattering function for the Wahnstr{\"o}m 50/50 binary LJ liquid at the wavevector corresponding to the static structure factor maximum. The red and black curves give, respectively, results for a FCS cutoff and large reference cutoff.}
\label{WaF}
\end{figure}
In summary, for all LJ type systems whenever there are strong virial potential-energy correlations ($R>0.9$), an FCS cutoff gives accurate results for both structure and dynamics. In the parts of the phase diagram where $WU$ correlations are weaker, an FCS cutoff gives poor results.
\subsection{Buckingham liquid}
Next, we consider the single-component Buckingham liquid ({SCB}). The Buckingham potential \cite{buckingham1,buckinghamisomorphs} is similar to the LJ potential, but does not have an IPL repulsive term; instead the potential's short-distance behavior follows a steep exponential (Fig. \ref{pairSCB}). The parameters of the Buckingham potential (Appendix \ref{pairP}) were chosen such that the LJ potential is well approximated in the repulsive region (Fig. \ref{pairSCB}).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/Buckingham_pair_potential}
\caption{The Buckingham pair potential (red) and the LJ pair potential (black). The parameters of the Buckingham potential were chosen such that the LJ potential is well approximated in the repulsive region.}
\label{pairSCB}
\end{figure}
Figures \ref{SCB}(a) and (b) show, respectively, the RDF and ISF for the SCB liquid. The FCS cutoff works well.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCB_rho1000_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SCB_rho1000_FsA}
\caption{The effect on structure and dynamics of varying the cutoff for the single-component Buckingham (SCB) liquid. The red and black curves give, respectively, results for an FCS cutoff and a large reference cutoff.
(a) RDF at $\rho=1.00$ and $T = 1.00$ ($R=0.99$). The inset quantifies the deviation in RDF from the reference RDF as a function of the cutoff.
(b) {ISF} at the same state point.}
\label{SCB}
\end{figure}
\subsection{Dzugutov liquid}
Figure \ref{pairDZ} shows the Dzugutov (DZ) pair potential \cite{dzugutov}, which was originally suggested as a model potential impeding crystallization by energetically punishing particle separations corresponding to the next-nearest neighbor distance of a face-centered cubic lattice. At short distances the {DZ} pair potential approximates the LJ potential.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/DZ_pair_potential}
\caption{The Dzugutov ({DZ}) pair potential \cite{dzugutov} (blue). Also shown is the LJ pair potential (black curve). The {DZ} potential approximates the LJ potential around the first minimum, but has a local maximum at larger distances.}
\label{pairDZ}
\end{figure}
Figures \ref{DZ}(a) and (b) show, respectively, the RDF and intermediate coherent scattering function of the DZ system. For this system the use of an FCS cutoff leads to poor results. This is not surprising given the fact that using an FCS cutoff removes the maximum of the {DZ} potential. What is important here, however, is that the poor FCS cutoff results correlate with the fairly weak virial/potential-energy correlations ($R=0.71$).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/DZ_rho0800_gr}
\includegraphics[width=70mm]{figs/compared/DZ_rho0800_F}
\vspace{0.5cm}
\caption{The effect on structure and dynamics of varying the cutoff for the Dzugutov (DZ) liquid at $\rho=0.80$ and $T = 0.75$ ($R=0.71$). The red and black curves give, respectively, results for an FCS cutoff and a large reference cutoff.
(a) RDF where the inset quantifies the deviation from the reference RDF as a function of the cutoff.
(b) The intermediate coherent scattering function at the same state point, including here results for a cutoff within the FCS (blue crosses).}
\label{DZ}
\end{figure}
\subsection{Lennard-Jones Gaussian liquid}
The Lennard-Jones Gaussian liquid \cite{LJG} (LJG) is a non-strongly correlating liquid with the two-minimum pair potential shown in Fig. \ref{pairLJG}. The parameters of LJG model (Appendix \ref{pairP}) are such that the second LJG potential minimum does not coincide with that of the {SCLJ} system \cite{ljgglassformer}.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/LJG_pair_potential}
\caption{The Lennard-Jones Gaussian (LJG) pair potential \cite{LJG} constructed by adding a Gaussian to the pair LJ potential. Two distinct minima are present.}
\label{pairLJG}
\end{figure}
Results from simulating structure and dynamics of the LJG liquid are shown in Figs. \ref{LJG}(a) and (b). The FCS cutoff does not give the correct RDF, whereas deviations in the dynamics are fairly small. Note that the FCS cutoff removes the second minimum.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/LJG_rho0850_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/LJG_rho0850_FsA}
\caption{The effect on structure and dynamics of the cutoff for the LJG liquid.
The red and black curves give, respectively, results for an FCS cutoff and a large reference cutoff.
(a) RDF at $\rho=0.85$ and $T = 2.00$ ($R=0.44$). The inset quantifies the deviation in RDF from the reference RDF (black curve) as a function of the cutoff.
(b) {ISF} at the same state point.}
\label{LJG}
\end{figure}
\subsection{Gaussian core model}
The Gaussian core model (GCM) \cite{gcm1,gcm2}, which is not strongly correlating, is defined by a Gaussian pair potential and thus has a finite potential energy at zero separation. The high-density regime of the {GCM} model ($\rho > 1.5$) has recently received attention as a single-component model glass former \cite{highdensitygcm} because it is not prone to crystallization and shows the characteristic features of glass-forming liquids (large viscosity, two-step relaxation, etc).
Figure \ref{GCM} shows the RDF and {ISF} for the {GCM} liquid. The {GCM} crystallizes when an FCS cutoff is used. For this reason, obviously, an FCS cutoff is not able to reproduce structure and dynamics of the reference system.
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/GCM_rho0850_rdf}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/GCM_rho0850_FsA}
\caption{The effect on structure and dynamics of varying the cutoff for the GCM liquid. The red and black curves give, respectively, results for an FCS cutoff and a large reference cutoff.
(a) RDF at $\rho=0.85$ and $T = 0.01$ ($R=-0.77$). The inset quantifies the deviation in RDF from the reference RDF as a function of the cutoff. The red curve represents a crystallized state.
(b) {ISF} at the same state points.}
\label{GCM}
\end{figure}
\subsection{The Hansen-McDonald molten salt model}
The final atomic system studied is the ``singly-charged molten salt model'' of Hansen and McDonald \cite{moltensalt}. In Fig. \ref{MS} we see that the structure is not represented well by the use of an FCS cutoff. Interestingly, the dynamics {\it is} well reproduced using this cutoff -- even better, in fact, than for a larger cutoff (Fig. \ref{MS}(b)).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/MS_rho0368_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/MS_rho0368_FsA}
\caption{The effect on structure and dynamics of varying the cutoff for Hansen-McDonald singly-charged molten salt model. Because of the competing interactions (Coulomb and $n=9$ repulsive IPL) this model is not strongly correlating. The red and black curves give, respectively, results for an FCS cutoff and a large reference cutoff.
(a) $AA$-particle RDF at $\rho=0.37$ and $T = 0.018$ ($R=0.15$). The inset quantifies the deviation in RDF from the reference RDF as a function of the cutoff.
(b) $A$-particle {ISF} at the same state point.}
\label{MS}
\end{figure}
\subsection{Two strongly correlating molecular model liquids}
We finish the presentation of the numerical results by giving data for three molecular model liquids. In this subsection data are given for two strongly correlating molecular liquid models, the Lewis-Wahnstr{\"o}m OTP \cite{otp1,otp2} and the asymmetric dumbbell \cite{sch09} models, which represent a molecule by three and two rigidly bonded LJ spheres, respectively. The next subsection gives data for a rigid water model.
Figures \ref{OTP}(a) and (b) show the LJ particle RDF and ISF of the OTP model. Both quantities are well approximated using an FCS cutoff, although slight deviations are noted for the ISF (red curve, see Appendix B for considerations concerning this). The OTP model is a border-line strongly correlating liquid ($R=0.91$).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/OTP_rho0986_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/OTP_rho0986_FsA}
\caption{The effect on structure and dynamics of varying the cutoff for the Wahnstr{\"o}m OTP model. The red and black curves give, respectively, results for an FCS cutoff and a large reference cutoff.
(a) RDF of the LJ particles at $\rho=0.33$ and $T = 0.70$ ($R=0.91$). The inset quantifies the deviation in RDF from the reference RDF as a function of the cutoff. The spikes derive from the bonds.
(b) {ISF} at the same state point.}
\label{OTP}
\end{figure}
Figures \ref{Asym}(a) and (b) show corresponding figures for the large (A) particle of the asymmetric dumbbell model at a viscous state point. The use of an FCS cutoff gives accurate results for both structure and dynamics. The FCS cutoff was placed at the {\it second} minimum of the AA RDF, because the {AA RDF} has here a lower value than at the first minimum. If the cutoff is placed at the first minimum, clear deviations are noted (data not shown).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/DB_rho1863_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/DB_rho1863_FsA}
\caption{The effect on structure and dynamics of varying the cutoff for the asymmetric dumbbell model.
The red and black curves give, respectively, results for an FCS cutoff and a large reference cutoff. Note that in this case the FCS cutoff is defined by using the {\it second} minimum (the first minimum is not the absolute minimum).
(a) RDF at $\rho=0.93$ and $T = 0.46$ ($R=0.96$). The inset quantifies the deviation in RDF from the reference RDF (black curve) as a function of the cutoff.
(b) A-particle {ISF} at the same state point.}
\label{Asym}
\end{figure}
\subsection{Rigid {SPC/E} water}
We consider finally the rigid {SPC/E} water model \cite{spce} (Fig. \ref{water}). This model is not strongly correlating at ambient conditions, a fact that reflects water's well-known density maximum \cite{paperII}. The structure of the {SPC/E} water model is not well represented using an FCS cutoff. The FCS cutoff dynamics on the other hand shows only slight deviations from that of the reference curve (black).
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SPCE_rho1000_gr}
\vspace{-20pt}
\end{figure}
\begin{figure}[H]
\centering
\includegraphics[width=70mm]{figs/compared/SPCE_rho1000_FsO}
\caption{The effect on structure and dynamics of varying the cutoff for the rigid {SPC/E} water model \cite{spce}.
The red and black curves give, respectively, results for an FCS and a large reference cutoff.
(a) Oxygen-oxygen RDF at $\rho=1.00$ and $T = 4.00$ ($R=0.08$). The inset quantifies the deviation in RDF from the reference RDF as a function of the cutoff.
(b) Oxygen {ISF} at the same state point.}
\label{water}
\end{figure}
\section{Summarizing the simulation results}\label{sum}
The previous section showed that structure and dynamics are well approximated in simulations using an FCS cutoff for the following atomic and molecular systems:
\begin{itemize}
\item Inverse power-law (IPL) systems ($n=18, 6, 4$),
\item single-component Lennard-Jones liquid at density $\rho=0.85$,
\item generalized Kob-Andersen binary Lennard-Jones mixtures,
\item Wahnstr{\"o}m binary Lennard-Jones mixture,
\item single-component Buckingham liquid,
\item Lewis-Wahnstr{\"o}m OTP model,
\item asymmetric dumbbell model.
\end{itemize}
These systems are all strongly correlating \cite{paperI,paperII,paperIII,paperIV,paperV}. Thus for strongly correlating liquids it is enough to know the intermolecular interactions within the FCS in order to accurately simulate structure and dynamics.
The simulations showed further that for all of the following atomic and molecular systems structure and/or dynamics are not properly reproduced when an FCS cutoff is used:
\begin{itemize}
\item Dzugutov (DZ) liquid,
\item Lennard-Jones Gaussian (LJG) liquid,
\item Gaussian core model (GCM),
\item Hansen-McDonald molten salt model,
\item rigid {SPC/E} water model.
\end{itemize}
None of these liquids are strongly correlating. For all these systems larger cutoffs are needed in order to faithfully reproduce the system's physics.
In conclusion, a shifted-forces FCS cutoff leads to accurate results {\it if and only if} the liquid is strongly correlating at the state point in question. We know of no exceptions to this rule. This suggests that strongly correlating liquids are {\it characterized} by the property that intermolecular interactions beyond the FCS can be ignored.
\section{The real essence of liquid simplicity}\label{SCL_sec}
As mentioned in the introduction a definition of simple liquids is most useful if it identifies their {\it real essence} \cite{locke}, the underlying fundamental characteristic from which these liquids' simple features, their {\it nominal essences}, follow. We suggest below that the class of simple liquids is to be identified with the class of strongly correlating liquids (Sec. \ref{sldef}). This is motivated by first summarizing the many simple properties of strongly correlating liquids (Sec. \ref{prop}), then showing that this class of liquids can be characterized from three different perspectives: mathematically, physically, and chemically (Sec. \ref{eqv}). This gives three very different but equivalent characterizations, indicating that the class of strongly correlating liquids is fundamental and further motivating the suggestion that the real essence of liquid simplicity is the existence of strong correlations of virial/potential-energy equilibrium $NVT$ fluctuations. By connecting to the chemists' concept of non-associated liquids we then discuss which real-world liquids are simple (Sec. \ref{simple}), discuss briefly liquids near interfaces (Sec. \ref{interface}), and give examples of complex liquid properties (Sec. \ref{complex}). Finally, Sec. \ref{perturbation} points out that our results call into question traditional perturbation theory, which is based on assuming quite different roles of the attractive and the repulsive forces.
\subsection{Strongly correlating liquids and their simple properties}\label{prop}
Most properties of strongly correlating liquids follow from the existence of ``isomorphs'' in their phase diagram (see below). Some of their properties were identified before isomorphs were defined in 2009 \cite{paperIV}, however, for instance that
\begin{itemize}
\item all eight fundamental thermoviscoelastic response functions are given in terms of just one, i.e., the dynamic Prigogine-Defay ratio is close to unity \cite{baileyOrderParameters},
\item aging may be described by adding merely one extra parameter \cite{paperII,paperIII},
\item power-law density scaling \cite{rol05} is obeyed to a good approximation, i.e., for varying density and temperature the relaxation time is a function of $\rho^\gamma/T$ \cite{sch09}.
\end{itemize}
An isomorph is an equivalence class of state points in a system's phase diagram. Two state points ($\rho_{1}$, $T_{1}$) and ($\rho_{2}$, $T_{2}$) are defined to be isomorphic if the following holds \cite{paperIV}: Whenever one microconfiguration of state point ($1$) and one of state point ($2$) have the same reduced coordinates (i.e., $\rho_{1}^{1/3} \textbf{r}^{(1)}_{i} = \rho_{2}^{1/3} \textbf{r}^{(2)}_{i}$ for all particles $i$), these two microconfigurations have proportional configurational Boltzmann factors,
\begin{equation} \label{defiso}
e^{-U(\textbf{r}_ {1}^{(1)}, ..., \textbf{r}_ {N}^{(1)})/k_{B}T_{1}} = C_{12}e^{-U(\textbf{r}_{1}^{(2)}, ..., \textbf{r}_ {N}^{(2)})/k_{B}T_{2}}.
\end{equation}
For most systems the isomorph concept is approximate just as $WU$ correlations are rarely perfect. Thus we do not require Eq. (\ref{defiso}) to be rigorously obeyed for all microconfigurations, but only to a good approximation and only for all {\it physically relevant} microconfigurations. By this is meant microconfigurations that are not {\it a priori} unimportant for the physics. Being an equivalence class, an isomorph defines a continuous curve of state points in the liquid's phase diagram. Only liquids for which the potential energy is an Euler homogeneous function, e.g., systems with IPL pair potentials, have exact isomorphs. An IPL fluid with $v(r)\propto r^{-n}$ has exact isomorphs characterized by $\rho^{\gamma} /T = {\rm Const.}$ where $\gamma = n/3$.
Appendix A of Ref. \onlinecite{paperIV} showed that a liquid is strongly correlating if and only if it has isomorphs to a good approximation. This was confirmed in Refs. \onlinecite{paperIV} and \onlinecite{paperV}, which showed that Lennard-Jones type atomic liquids have good isomorphs. Reference \onlinecite{ing11} showed that the strongly correlating Lewis-Wahnstr{\"o}m OTP and asymmetric dumbbell molecular models also have good isomorphs.
Equation (\ref{defiso}) has many consequences. These were derived and discussed in detail in the original isomorph paper from 2009 (Ref. \onlinecite{paperIV}), to which the reader is referred. Basically, structure and dynamics at two isomorphic state points are identical in reduced units. Quantities that are invariant along an isomorph include (but are not limited to):
\begin{enumerate}
\item The excess entropy, i.e., the entropy in excess of the ideal gas entropy at the same density and temperature -- this is the configurational contribution to the entropy (a term that is negative because a liquid is always more ordered than an ideal gas at same density and temperature).
\item All $N$-body entropy terms. Recall that the excess entropy can be expanded in a series of two-body, three-body, etc, terms. Each term is invariant along an isomorph \cite{paperIV}.
\item The isochoric heat capacity.
\item The structure in reduced units (defined by $\tilde{\textbf{r}}_{i} \equiv \rho^{1/3}\textbf{r}_{i}$ for all particles $i$). Not only the radial distribution function, but all higher-order distribution functions in reduced units are isomorph invariant.
\item The Newtonian {\it NVE} and Nos$\acute{e}$-Hoover {\it NVT} equations of motion in reduced units; likewise Brownian dynamics.
\item All autocorrelation functions in reduced units.
\item All average relaxation times in reduced units.
\item Reduced transport coefficients like the diffusion coefficient, viscosity, etc.
\end{enumerate}
Isomorphs have the further interesting property that there is no relaxation for an instantaneous change of temperature and density when jumping from an equilibrated state point to a different state point isomorphic with the initial state. The absence of relaxation derives from the fact that the Boltzmann probabilities of scaled microconfigurations are identical. Such ``isomorph jumps'' work quite well for the KABLJ liquid \cite{paperIV}, for the asymmetric dumbbell, and for the Lewis-Wahnstr{\"o}m OTP molecular models \cite{ing11}. Moreover, the effective temperature of a glass prepared by a temperature-density jump from an equilibrium state of a strongly correlating liquid depends only on the final density \cite{effectivetemperature}; this provides yet another example of a simple feature of these liquids.
Some further predictions for the class of strongly correlating liquids deriving from the existence of isomorphs are:
\begin{itemize}
\item The solid-liquid coexistence curve is an isomorph \cite{paperIV,paperV}. This implies invariance along the coexistence curve of the reduced structure factor, the reduced viscosity, the reduced diffusion constant, etc, as well as pressure invariance of the melting entropy, and a reduced-unit Lindemann melting criterion \cite{paperIV}.
\item Collapse of the two-order-parameter maps of Debenedetti {\it et al.} \cite{order1,order2,order3,order4,order5} to one-dimensional curves \cite{paperIV}.
\item Isochronal superposition \cite{nga05}, i.e., the fact that when pressure and temperature are varied, the average relaxation time determines the entire relaxation spectrum \cite{paperIV}.
\end{itemize}
The above listed properties of strongly correlating liquids all reflect {\it simple} features of strongly correlating liquids. A final, recently established simple property is a thermodynamic separation identity: for all strongly correlating liquids, if $s$ is the excess entropy per particle, the temperature as a function of $s$ and density $\rho$ factorizes as follows \cite{ing11a}
\begin{equation}\label{eos}
T\,=\,f(s)h(\rho)\,.
\end{equation}
Equation (\ref{eos}) has a number of consequences \cite{ing11a}, including the Gruneisen equation of state and that the isomorphs of LJ liquids -- in particular, the LJ solid-liquid coexistence curve -- are given by $(A\rho^4-B\rho^2)/T={\rm Const.}$ \cite{khrapak,beyond}.
\subsection{Mathematical, physical, and chemical characterizations of strongly correlating liquids}\label{eqv}
At a given state point, if the average potential energy is denoted by $\langle U\rangle$, the constant-potential-energy hypersurface is defined by $\Omega=\{(\br_1,...,\br_N)\in R^{3N}|U(\br_1,...,\br_N)=\langle U\rangle\}$. This is a compact, Riemannian $(3N-1)$-dimensional differentiable manifold. Each state point has its own such hypersurface. In this way, a family of manifolds is defined throughout the phase diagram. In Appendix A of Ref. \onlinecite{paperIV} it was shown that the reduced-unit constant-potential-energy manifold is invariant along a strongly correlating liquid's isomorphs, and that invariance curves exist for these manifolds only for strongly correlating liquids. Thus, for such liquids these manifolds constitute a one-parameter family of manifolds, not two-parameter families as expected from the fact that the phase diagram is two-dimensional.
The above provides a {\it mathematical} characterization of the class of strongly correlating liquids. The {\it physical} characterization was discussed already: the existence of isomorphs in the phase diagram. A liquid is strongly correlating if and only if it has isomorphs to a good approximation (Appendix A of Ref. \onlinecite{paperIV}).
The {\it chemical} characterization of strongly correlating liquids is the property documented in the present paper: A liquid is strongly correlating at a given state point if and only if the liquid's structure and dynamics are accurately calculated by shifted-forces cutoff simulations that ignore interactions beyond the first coordination shell. This is an empirical finding for which we have at present no compelling arguments, but can at present merely point out two things. First, the property of insignificance of interactions beyond the FCS is an isomorph invariant: {\it If} a liquid has good isomorphs and {\it if} an FCS cutoff works well at one state point, FCS cutoffs must work well for all its isomorphic state points. Thus the chemical characterization of strongly correlating liquids is consistent with the fact that these liquids have isomorphs. Secondly, almost all of the fluctuations in virial and potential energy of the LJ liquid come from interparticle separations within the FCS \cite{paperII}.
The new ``FCS characterization'' of strongly correlating liquids shows that these liquids are characterized by having a well-defined FCS. Most likely it is the existence of a well-defined FCS which implies the almost cancellation of the linear term of the shifted-force potential. The fact that interactions beyond the FCS may be ignored shows that interactions are effectively short ranged, which means that the structure is dominated by what may be termed packing effects.
\subsection{Defining the class of ``simple liquids''}\label{sldef}
Section \ref{prop} listed strongly correlating liquids' many simple properties. Section \ref{eqv} showed that this liquid class may be characterized from three quite different points of view. Clearly the class of strongly correlating liquids is fundamental. Since the properties of strongly correlating liquids are generally simpler than those of liquids in general, we now propose the definition
\begin{itemize}
\item {\bf Simple liquids = Strongly correlating liquids}
\end{itemize}
This is the basic message of the present paper, which implies a quantification of the degree of simplicity via the number $R$ of Eq. (\ref{Rdef}), the $NVT$ ensemble equilibrium virial/potential-energy correlation coefficient.
Compared to the standard definition of simple liquids (as those with radially symmetric pair interactions) there are some notable differences:
\begin{enumerate}
\item Simplicity is quantified by a continuously variable, it is not an on/off property.
\item The degree of simplicity generally varies throughout the phase diagram. Most strongly correlating liquids loose this property as the critical point is approached; on the other hand many or most non-strongly correlating liquids are expected to become strongly correlating at very high pressures \cite{highpressurescl}.
\item Not all ``atomic'' liquids (i.e., with radially symmetric pair interactions) have simple regions in the low-pressure part of the phase diagram (compare the Dzugutov, Lennard-Jones Gaussian, Gaussian core, and molten salt models).
\item Not all simple liquids are atomic (compare the Wahnstr{\"o}m OTP and the asymmetric dumbbell models).
\end{enumerate}
According to the new definition of liquid simplicity the case where the potential energy is an Euler homogeneous function of the particle positions ($R=1$) sets the gold standard for simplicity. This is consistent with the many simple properties of IPL liquids. Due to the absence of attractions, IPL fluids have no liquid-gas phase transition. In this sense it may seem strange to claim that IPL fluids are the simplest liquids. However, more realistic strongly correlating liquids like the LJ liquid cease to be so when the liquid-vapor coexistence line is approached near the critical point, showing that this phase transition cannot be understood in the framework of simple liquids. This contrasts with the liquid-solid phase transition, where for instance the fact that the coexistence line for simple liquids is an isomorph -- confirmed by simulations of the LJ liquid \cite{paperV} -- explains several previously noted regularities \cite{paperIV,khrapak}.
Is the hard-sphere fluid simple? One may define a configurational virial function for this system, but it is not obvious how to define a potential energy function that is different from zero. Thus there is no meaningful correlation coefficient $R$ for hard-sphere fluids. On the other hand, the hard-sphere liquid may be regarded as the $n\rightarrow\infty$ limit of an IPL liquid, and it is well known that for instance the hard-sphere radial distribution function is close to that of, e.g., an $r^{-20}$ IPL liquid at a suitably chosen temperature. This would indicate that hard-sphere liquids are simple, consistent with the prevailing point of view. Another interesting case is that of the Weeks-Chandler-Andersen (WCA) version of the LJ liquid, which cuts off all attractions by putting the force equal to zero beyond the potential energy minimum. This liquid is strongly correlating \cite{cos09}, but in simulations that the WCALJ liquid has somewhat poorer isomorphs than the LJ liquid.
It is possible that the hard-sphere liquid and the WCALJ liquid should be both excluded from the class of simple liquids on the grounds that their potentials are not analytic. For systems interacting with pair potentials, for instance, it could make good sense to add the extra requirement that the pair potential is an analytical function of the inverse pair distance, i.e., that an expansion exists of the form $v(r)=\sum_n v_n r^{-n}$. Such an extra analyticity requirement would not exclude any strongly correlating liquids occurring in nature where all potentials are expected to be analytic.
\subsection{Which liquids in the real world are simple?}\label{simple}
Real-world liquids may be classified according to the nature of the chemical bonds between the molecules. There are five types of bonds \cite{pauling}, listed below with a few typical examples (polymeric systems may be added as a separate class):
\begin{itemize}
\item Van der Waals bonds (argon, toluene, butane, ...);
\item Metallic bonds (gold, aluminum, alloys, ...);
\item Hydrogen bonds (water, glycerol, ethanol, ...);
\item Ionic bonds (molten sodium chloride and potassium nitrate, room-temperature ionic liquids, ...);
\item Covalent bonds (silica and borate melts, ...).
\end{itemize}
Most liquids involve elements of more than one type of chemical bonds. For instance, van der Waals forces are present in all liquids; the first class consists merely of those liquids that {\rm only} have van der Waals forces. Another borderline example is a dipolar organic liquid like di-butyl-phthalate, where van der Waals as well as Coulomb forces are present; a liquid like glycerol has also strong dipolar interactions, i.e., an element of the ionic bonds that defines class 4), etc.
Based on computer simulations and known properties of liquids we believe that most or all van der Waals and metallic liquids are strongly correlating \cite{ped08,baileyOrderParameters,paperII}, i.e., {\it simple}. Liquids that are not simple are the hydrogen-, ionically, and covalently bonding liquids. In these cases the virial/potential-energy correlations are weakened by one or the other form of competing interactions.
Metals play a special role as simple liquids, because their interatomic forces derive from collective interactions between ion cores and free electrons \cite{fis64}. The resulting interaction is a non-directional interaction between symmetric ion cores, i.e., these systems are ``simple'' in the traditional sense. Preliminary computer simulations of ours show that metals are strongly correlating \cite{paperI}, so metals are simple also in the sense of this paper, as well. However, not all isomorph invariants apply for metals. For instance, the electron gas can influence the collective dynamics without any visible structural and relaxational counterpart \cite{bov01,pet11}, so isomorph invariance most likely breaks down for these (fast) collective degrees of freedom.
It should be emphasized that the above considerations refer to ambient or moderate pressure conditions. It was recently suggested that if crystallization is avoided, all liquids become strongly correlating at sufficiently high pressure \cite{highpressurescl}. Thus, e.g., the molten silicates of the earth's upper mantle are predicted to be simpler than molten silicates at ambient pressure.
\subsection{Liquids near interfaces}\label{interface}
Since the property of being simple cannot be read of from knowledge of the potential alone, it is of interest to consider liquids under more general circumstances, for instance under confinement or generally near interfaces. Liquids near interfaces show rich and complicated behavior. For instance, a liquid confined to the nanoscale may change its dynamic properties several orders of magnitude compared to the bulk system. Predicting these changes is an important challenge relevant for biological systems, engineered devices, etc. Recently, it was shown that some liquids retain bulk liquid behavior in confinement
\cite{mit06,LJconfinedsmoothwalls,connectiondynamicsprofile,dumbbellconfinedroughwalls}. More specifically, it was shown that Rosenfeld's excess entropy scaling in the bulk persists in confinement and is to a good approximation independent of the wall-fluid interaction strength and the degree of confinement. This was shown for LJ and hard-sphere liquids, suggesting the possibility of extending the concept of a simple liquid to apply beyond bulk systems. Given that interactions in strongly correlating liquids are limited in range by the radius of the FCS, one may speculate that these liquids are simple also by having the property that an external field at one point affects the liquid over shorter distances than for a general liquid. Clearly, more work is needed to clarify the relevance and consequences of the present definition of liquid simplicity near interfaces and in external fields \cite{wee97,wee02}.
\subsection{A note on complex liquid behavior}\label{complex}
Liquids that are not simple in the above defined sense of the term often have complex properties \cite{order3,SCS1,SCS2,SCS3,tetraAnomalies}. Water with its correlation coefficient close to zero at ambient conditions is a prime example of a complex liquid. It is well known for water that a certain region of state points in the density/temperature phase diagram exhibits anomalous thermodynamic behavior in the sense that isobaric heating implies densification. Numerical evidence indicates that these state points lie within a larger region with diffusion anomaly, i.e., an increased diffusivity upon isothermal compression \cite{order3}, a region that in turn lies within a larger region of structural anomaly characterized by decreasing order upon isothermal compression \cite{order3}.
Different order parameters exist for characterizing the structural order of liquids, some of which relate purely to an integral over the RDF \cite{SCS1,SCS2,SCS3}. In this way it is possible to calculate the contribution to structural anomalies from the different coordination shells \cite{SCS1,SCS2,SCS3}. It has been shown \cite{SCS2,SCS3} that the structural anomaly of water and waterlike liquids is not a: ''{... first-shell effect. Rather, they reflect how structuring in second and more distant coordination shells responds to changes in thermodynamic or system parameters.}'' The full anomalous behavior of water derives from interactions beyond the FCS \cite{SCS1,SCS3}. This is consistent with the results presented in this paper, since the structure and dynamics of strongly correlating liquids are given exclusively by the interactions within the FCS.
\subsection{To which extent do the assumptions of standard pertubation theory hold?}\label{perturbation}
The finding that the FCS plays a crucial role for a large class of systems may be taken as a modern demonstration of the classic van der Waals picture of liquids in the sense that such liquids can be understood in terms of packing effects \cite{row82}. On the other hand, our results call into question the basis of traditional perturbation theory, which is also usually traced back to van der Waals \cite{wid67}. Perturbation theory is based on the assumption of entirely different roles being played by the repulsive and the attractive forces \cite{wid67,wca,bar76,han05,row82,bar76,wee97,zho09}: The repulsive forces define the structure and reduce the entropy compared to that of an ideal gas at same density and temperature, the attractive forces reduce the pressure and energy compared to that of an ideal gas. From the findings of the present and a previous paper \cite{FCS2} it is clear, however, that this picture applies only when the FCS coincides roughly with the pair potential minimum, i.e., at fairly low pressure. At very high pressures the whole FCS is within the range of the repulsive forces, here the attractive forces play little role for simple liquids. In general, what is important for a strongly correlating liquid is to take into account properly all forces from particles within the FCS -- and only these. Thus the well-known WCA reference system, which ignores the attractions, is a good reference only at such high pressures that all forces from particles within the FCS are repulsive \cite{wcabertier1,wcabertier2,FCS2}.
The dominance of the FCS for simple liquids reflects the fundamental physics that the characteristic length defining the potential minimum (e.g., $\sigma$ of the LJ potential) is much less important than generally believed: $\sigma$ determines the density of the low-pressure condensed phase, but that is all. The only physically important length for simple liquids is that given by the macroscopic density itself: $\rho^{-1/3}$. At low pressures this length is roughly that of the potential energy minimum, explaining why the latter has been generally assumed to be important.
The above considerations apply only for simple liquids; in general both lengths play important roles for the physics. The irrelevance of any length defined by the microscopic potential emphasizes once again that the class of strongly correlating liquids is at the one end of the ``complexity scale'' where, at the other end, one finds systems like macromolecules, electrolytes, interfaces, micelles, or enzymes, for which multiple length and time scales are important \cite{bag10}.
\section{Concluding remarks}\label{disc}
If you ask a chemist what is a simple liquid, he or she may answer that nonassociated liquids are simple, whereas associated liquids are much more complex. These two concepts are defined as follows in Chandler's textbook \cite{chandler}. The intermolecular structure of a {\it nonassociated} liquid ``can be understood in terms of packing. There are no highly specific interactions in these systems.'' In contrast, water is an example of an {\it associated} liquid, and its ``linear hydrogen bonding tends to produce a local tetrahedral ordering that is distinct from what would be predicted by only considering the size and shape of the molecule'' \cite{chandler}.
Packing usually refers to purely entropic, hard-sphere like behavior. Given that no realistic potentials are infinitely repulsive it makes good sense to interpret packing more generally as all short-ranged effects of the intermolecular interactions. If one accepts this more general interpretation of packing, the crucial role of the FCS for strongly correlating liquids is consistent with the understanding since long that the properties of nonassociated liquids can be interpreted in terms of packing:
\begin{itemize}
\item Once the forces from particles within the FCS are known, basically everything is known.
\end{itemize}
In other words, for a simple liquid there are no important long-range interactions, and ``considering the size and shape of the molecule'' is enough to account for the liquid's physical properties. This applies even for the $r^{-4}$ IPL fluid, which has a fairly long-ranged interaction.
The present definition of the class of simple liquids is thus consistent with the chemists' general picture of simple liquids. The new definition goes further, however, by quantifying simplicity via the virial/potential-energy correlation coefficient $R$ of Eq. (\ref{Rdef}). In particular, the degree of simplicity is not an on/off property of the potential, but varies continuously with state point. Thus even a complex liquid like water is expected to approach simple behavior under sufficiently high pressure \cite{highpressurescl} and, conversely, the prototype strongly correlating LJ liquid becomes gradually more complex as density is lowered and the critical region and/or the gas phase is approached. Is this a problem, given that everyone agrees that the gas phase is simple? We do not think so. In fact, the gas phase is simple for a different reason, namely that molecules move freely most of the time, only interrupted by occasional fast and violent collisions with other molecules. It would be strange if a system exhibiting one form of simplicity could be transformed continuously, while maintaining its simplicity, into a system of an entirely different form of simplicity; one would expect the intermediate phase to be complicated.
Liquid simplicity is characterized by the correlation coefficient $R$ of Eq. (\ref{Rdef}) being close to unity, i.e., that $1-R$ is a small number. This situation is typical in physics, where simplifying features always appear when some dimensionless number is small. The obvious question arises whether a perturbation theory may be constructed around simple liquids, embracing the more complex ones. Only future can tell whether this is possible, but it does present a challenge because the properties of IPL fluids ($R=1$) cannot be worked out analytically.
A potentially annoying feature about defining liquid simplicity from the existence of strong correlations of the virial/potential-energy fluctuations is that one cannot read off directly from potential and state point whether or not a given liquid is simple. The same holds for the existence of isomorphs or the property that an FCS cutoff reproduces the correct physics. We believe one should accept this as an acceptable cost for precisely delimiting the simple liquids from others. With the power of today's computers this is much less of a problem than previously. For most systems a brief simulation will determine whether or not the liquid is strongly correlating at the state point in question.
Except for the IPL fluids, no system is simple in the entire fluid phase. This paper focused on the condensed liquid phase, not too far from the solid-liquid coexistence line, but far from the critical point and the gas phase -- it is here that some liquids are simple. The focus on liquids is not meant to imply a limitation to the liquid phase, however. In fact, simulations show that when a strongly correlating liquid crystallizes, the crystal is at least as strongly correlating \cite{paperII}. A theory has been developed for (classical) strongly correlating crystals, showing that the property of strong virial/potential-energy equilibrium fluctuations in the $NVT$ ensemble is an anharmonic effect that survives as $T\rightarrow 0$ \cite{paperII}. Of course, low-temperature crystals are not classical systems, and both for liquids and crystals an interesting topic for future works is the implication of the proposed simplicity definition for the quantum description of condensed matter.
Section \ref{prop} summarized the several nominal essences of simple liquids. What is the {\it real essence} of liquid simplicity? Given that three fundamental characterizations of strongly correlating liquids are equivalent -- the mathematical, the physical, and the new chemical (FCS) characterization -- this question cannot be answered unequivocally. At the end of the day it is a matter of taste whether one defines liquid simplicity from the existence of strong virial/potential-energy correlations, from the existence of isomorphs, from the existence of invariance curves for constant-potential-energy hypersurfaces, or from the property that interactions beyond the FCS play little role. All four properties are equivalent.
\acknowledgments
The centre for viscous liquid dynamics ``Glass and Time'' is sponsored by the Danish National Research Foundation (DNRF). We gratefully acknowledge useful inputs from Livia Bove, Jesper Schmidt Hansen, and S{\o}ren Toxv{\ae}rd.
|
\section{Introduction and summary}
\label{sec:Introduction}
Wilson loops and their magnetic cousins, 't Hooft loops,
are universal observables in gauge theories
whose properties characterize the phases of each theory.
They represent heavy probe particles with electric and magnetic charges
moving along a closed trajectory in spacetime.
When acting on the Hilbert space, these operators do not commute
if the two loops are linked within the constant time slice
\cite{'tHooft:1977hy}.
Indeed 't Hooft successfully used their non-trivial commutation relations
to classify the possible phases of non-Abelian gauge theories
\cite{'tHooft:1977hy,'tHooft:1979uj,tHooft:1981ht}.
Noncommutativity is also a hallmark of quantization.
The position and the momentum of a particle do not commute with each other;
they cannot be simultaneously diagonalized or precisely measured.
For physicists quantization is usually the process of obtaining
a Hilbert space and noncommuting operators acting on it from a classical system.
In certain situations (especially for mathematicians), however, one is
primarily interested in the ``operators'' without a Hilbert space.
In such a scheme, called deformation quantization,
the product of two functions on a phase space (Poisson manifold)
is continuously (in $\hbar$) deformed into a noncommutative associative product
whose order $\mathcal O(\hbar)$ correction is given by the Poisson bracket.
It is a non-trivial result that
any Poisson structure admits a canonical deformation quantization \cite{Kontsevich:1997vb}.
In this paper we study Wilson-'t Hooft line operators in
$\mathcal N=2$ supersymmetric gauge theories on $S^1\times\mathbb R^3$
in the Coulomb phase.
We consider half-BPS line operators
\cite{Kapustin:2005py} extended along $S^1$
and perform an exact localization calculation of their expectation values
(vevs) following \cite{Pestun:2007rz}.
The vev of the product of operators turns out to
be given by the Moyal product of the vevs of the individual operators.
For an $\mathcal N=2$ theory characterized by a punctured Riemann surface
\cite{Gaiotto:2009we}, the line operators precisely realize
the deformation quantization of the Hitchin moduli space, with respect to the
Poisson structure specified by the complexified Fenchel-Nielsen coordinates
\cite{Nekrasov:2011bc,Dimofte:2011jd}.
Let us summarize our main results in more detail.
The vev of a line operator is a holomorphic function
of $a$ and $b$, which take values
in the complexified Cartan subalgebra $\mathfrak t_{\mathbb C}$
and its dual $\mathfrak t^*_{\mathbb C}$.
The variable $a$ is a combination
of the electric Wilson line $A_\tau$ and a real vector multiplet scalar,
while $b$ combines the magnetic Wilson line
and the other real scalar in the vector multiplet, all evaluated at
the infinity of $\mathbb R^3$.
These variables $a$ and $b$ parametrize the Coulomb branch of the gauge theory compactified
on $S^1$.
Since the path integral for the vev defines
a supersymmetric index (\ref{trace-def}), the electric and magnetic Wilson lines
can be regarded as chemical potentials for electric and magnetic charges.
The vev also depends holomorphically on the non-dynamical variables
$m_f$, which are complex combinations of masses
and chemical potentials for flavor symmetries.
Importantly, a real parameter $\lambda$ also enters in the vev.
It is defined as the chemical potential for the simultaneous
spatial and R-symmetry rotations.
We find that the vev of the Wilson operator in representation $R$
is simply given by
\begin{equation}
\langle W_R\rangle= {\rm Tr}_R\, e^{2\pi i a}\,,
\end{equation}
where the trace is taken in $R$.
The localization calculation is only non-trivial for 't Hooft and dyonic
line operators.
In particular, they have a non-trivial one-loop determinant as well
as the non-perturbative contributions from Polyakov-'t Hooft monopoles
screening the charges of the singular monopole \cite{Kapustin:2006pk,Gomis:2011pf}.
The 't Hooft operator specified by a coweight $B$ has a vev of the form
\begin{equation}
\langle T_B\rangle=
\sum_v e^{2\pi i v\cdot b}
Z_\text{1-loop}(a,m_f,\lambda;v)
Z_\text{mono}(a,m_f,\lambda;B,v)\,.
\label{eq:intro-thooft}
\end{equation}
For simplicity we often suppress the dependence on some of $a$, $m_f$ and $\lambda$.
The function $Z_\text{1-loop}(v=B)$ given in (\ref{1-loop-vm}-\ref{1-loop-total})
is the one-loop determinant
around the leading saddle point.
The sum is over the magnetic charges $v$ reduced from $B$ due to monopole screening.
There exists a number of non-perturbative
saddle points that correspond to coweight $v$,
and $Z_\text{1-loop}(v)Z_\text{mono}(B,v)$
is the sum of the fluctuation determinants around the saddle points
in the sector $v$.
The function $Z_\text{mono}(B,v)$, given in
(\ref{mono-adj}, \ref{mono-fund}, \ref{mono-total})
for $G=U(N)$ and matter in the adjoint or fundamental representation,
is a monopole analog of the Nekrasov
instanton partition function \cite{Nekrasov:2002qd}.
For a more general dyonic Wilson-'t Hooft operator,
we insert into (\ref{eq:intro-thooft})
a Wilson operator in the subgroup of the gauge group unbroken by $B$.
Let us suppose that the spatial rotation associated with $\lambda$
takes place in the 12-plane.
It is useful to think of the 3-axis as the Euclidean time direction,
and consider line operators $L_i$ at various points on the axis.
By the original argument of 't Hooft, we show that
these operators form a noncommutative algebra, generalizing the standard
't Hooft commutation relation $W\cdot T=e^{2\pi i/N} T\cdot W$
for minimal Wilson ($W$) and 't Hooft ($T$) loops in $SU(N)$ gauge theories.\footnote{Non-commutativity can also be understood by thinking of $S^1$ as time.
The electric and magnetic fields produced by the Wilson and 't Hooft operators
generate a non-zero Poyinting vector carrying angular momentum,
and contribute non-trivially to the supersymmetric index with $\lambda \neq 0$
\cite{Gaiotto:2010be}.
}
Moreover, operator multiplication is implemented by
noncommutative associative products, i.e.,
\begin{equation}
\langle L_1\cdot L_2\cdot\ldots \cdot L_n\rangle
= \langle L_1\rangle *\langle L_2\rangle *
\ldots *
\langle L_n\rangle\,,
\end{equation}
where
\begin{equation}
(f*g)(a,b)\equiv
\left.
e^{i\frac{\lambda}{4\pi}
(
\partial_{b}\cdot \partial_{a'}
-
\partial_a\cdot \partial_{b'}
)
}f(a,b) g(a',b')
\right|_{a'=a,b'=b}\,.
\end{equation}
This is the Moyal product associated with the
Poisson structure determined by the holomorphic symplectic form
\begin{equation}
da\wedge db
\label{eq:sympl-intro}
\end{equation}
and $\hbar=\frac{\lambda}{2\pi}$,
where $a$ and $b$ are contracted in the canonical way.
It is illuminating to be even more explicit, focusing
on the $SU(2)$ $\mathcal N=2^*$ theory.
In this case the expectation values of the minimal
Wilson ($W$), 't Hooft ($T$), and dyonic ($D$) operators are given by\footnote{For $SU(2)$ we simplify notation by substituting
$a\rightarrow \text{diag}(a,-a),
b\rightarrow \text{diag}(b,-b)$.
When there is only one mass parameter, $m\equiv m_{f=1}$.
}
\begin{equation}
\begin{aligned}
& \langle W\rangle=\, e^{2\pi i a}+e^{-2\pi i a}\,,
\quad\langle T\rangle=
\left( e^{2\pi i b}+e^{-2\pi i b}\right)
\left(
\frac{
\sin\left(2\pi a+\pi m \right)
\sin\left(2\pi a-\pi m\right)
}{
\sin\left(2\pi a+\frac{\pi}2 \lambda\right)
\sin\left(2\pi a-\frac{\pi}2 \lambda\right)
}
\right)^{1/2}\hspace{-3mm}
\,,
\\
&\hspace{20mm}
\langle D\rangle
=
\left( e^{2\pi i (b+a)}+e^{-2\pi i (b+a)}\right)
\left(
\frac{
\sin\left(2\pi a+\pi m \right)
\sin\left(2\pi a-\pi m\right)
}{
\sin\left(2\pi a+\frac{\pi}2 \lambda\right)
\sin\left(2\pi a-\frac{\pi}2 \lambda\right)
}
\right)^{1/2}\,.
\end{aligned}
\end{equation}
For $\lambda=0$, these expressions precisely appeared in \cite{Nekrasov:2011bc}
as the definition of Darboux coordinates $(a,b)$ on the Hitchin
moduli space for a one-punctured torus.
In \cite{Dimofte:2011jd},
they
were identified
as the complexification of the Fenchel-Nielsen coordinates,
which are Darboux coordinates for the real symplectic structure
on Teichm\"uller space.
Their findings are consistent with our identification
(\ref{eq:sympl-intro}) of the symplectic structure.
For $\lambda\neq 0$, our results provide quantum deformations.
Thus $\mathcal N=2$ gauge theories on $S^1\times \mathbb R^3$
produce a noncommutative algebra of operators quantizing the Hitchin moduli space.
Is there a Hilbert space on which the line operators
naturally act?
We claim that the space of conformal blocks in Liouville or Toda conformal
field theories
is such a Hilbert space.
This is demonstrated by showing that the Verlinde operators\footnote{Verlinde operators are the difference operators that act on conformal blocks, and
arise from the monodromy of extended conformal blocks with degenerate field insertions.
Verlinde operators in Liouville theory coincide with the geodesic length operators in quantum Teichm\"uller theory \cite{Teschner:2005bz}.
}
\cite{Drukker:2009id,Alday:2009fs},
labeled by closed curves on the Riemann surface
and corresponding to line operators in gauge theories \cite{Drukker:2009tz},
are exactly the Weyl transform (also known as the Weyl ordering) of
the vevs of the line operators on $S^1\times \mathbb R^3$,
where $a$ and $b$ are treated as coordinates and momenta, respectively.
The twist/quantization variable $\lambda$ is related to the variable
${\mathsf b}$ that parametrizes the central charge $c=1+6({\mathsf b}+{\mathsf b}^{-1})^2$
as
$ \lambda={\mathsf b}^2 $.
This result is a concrete realization of the proposal
that the algebra of line operators provide quantization
of the Hitchin moduli space
\cite{Nekrasov:2010ka,Teschner:2010je,Gaiotto:2010be,Dimofte:2011jd}.
The connection to Liouville/Toda theories provides a very strong
check of our localization computations.
Moreover, we conjecture that the connection should hold
even when $\mathcal N=2$ gauge theories have no Lagrangian description.
Thus it is now possible to compute the line operator vevs
on $S^1\times \mathbb R^3$ for such theories as
the inverse Weyl transform of the Verlinde operators.
Then the AGT relation \cite{AGT} between Liouville/Toda theories
and four-dimensional gauge theories
would suggest that our analysis should intimately parallel
the localization computation of 't Hooft loops on $S^4$
\cite{Gomis:2011pf} corresponding to $\lambda={\mathsf b}^2=1$.
Indeed $Z_\text{1-loop}(\lambda;B)$ and $Z_\text{mono}(\lambda;B,v)$
with $\lambda=1$ appeared in
\cite{Gomis:2011pf} as the contributions from the equator $S^1$ of $S^4$,
where a 't Hooft loop was inserted.
Exactly the same physical system on $S^1\times \mathbb R^3$
was considered
in \cite{Gaiotto:2010be},
where supersymmetric line operators were
analyzed from the point of view of wall-crossing in the IR effective theories.
Their twist parameter $y$ is given by $y=-e^{\pi i\lambda}$.
Based on the consistency of the wall-crossing formula
in $\mathcal N=2$ gauge theories and several other assumptions,
they conjectured expressions for the line operator vevs
in terms of the (commutative and noncommutative) Fock-Goncharov coordinates on
the Hitchin moduli space.
It would be desirable to perform more detailed comparisons.
This should help create a bridge between the AGT correspondence
\cite{AGT} and the study of wall-crossing, perhaps along the line of
\cite{Dimofte:2011ju}.
This paper is organized as follows.
Section \ref{sec:set-up} defines the gauge theory setup
and the quantities we wish to compute.
We begin our localization calculations in Section \ref{sec:localization},
where we analyze the symmetries of the system and lay out our strategy.
We also calculate the classical on-shell action
in the supersymmetric background defining a 't Hooft operator.
Section \ref{sec:one-loop} is devoted to the one-loop analysis.
In Section \ref{sec:mono} we compute the non-perturbative contributions due
to monopole screening.
Putting all together the classical, one-loop, and screening contributions,
Section \ref{sec:gauge-results} summarizes the results of our localization calculations and gives explicit expressions in several examples.
We then turn to the quantization aspects of our results.
In Section \ref{sec:non-com} we study the noncommutative structure
in the algebra formed by line operators and show that it implements
the deformation quantization of the Hitchin moduli space.
Next we discuss the relation to gauge theories on $S^4$ and Liouville/Toda
theories in Section \ref{sec:S4-CFT}.
We conclude the paper in Section \ref{sec:discussion}
with a discussion on related works and future directions.
Appendix \ref{sec:spinor-gamma} explains our convention
for spinors and gamma matrices.
In Appendix \ref{sec:mono-inst} we review Kronheimer's correspondence
between singular monopoles and $U(1)$-invariant instantons on a Taub-NUT space.
This relation is used in Sections \ref{sec:one-loop}
and \ref{sec:mono}.
Appendix \ref{sec:CFT-details} contains technical computations in Liouville and
Toda theories.
In Appendix \ref{sec:cl-holo}
we compute the classical $SL(2,\mathbb C)$ holonomies
on the four-punctured sphere and compare them with
gauge and Liouville calculations.
\section{$\mathcal N=2$ gauge
theories on $S^1\times \mathbb R^3$
and line operators}
\label{sec:set-up}
In this paper we study four-dimensional gauge theories with
$\mathcal N=2$ supersymmetry on $S^1\times \mathbb R^3$
in the Coulomb branch.
For notational convenience, we will use the
notation appropriate for $\mathcal N=2^*$ theory,
which can be thought of as a dimensional reduction of
the ten-dimensional super Yang-Mills,
though we will state
general results applicable to other field contents
\cite{Pestun:2007rz,Gomis:2011pf}.
The ten-dimensional gauge field $A_M$ ($M=1,\ldots, 9,0$) gives rise
to the four-dimensional gauge field $A_\mu$ ($\mu=1, \ldots, 4$),
hypermultiplet scalars $A_i\equiv \Phi_i$ ($i=5,\ldots, 8$),
and vector multiplet scalars $A_A\equiv \Phi_A$ ($A=0,9$).
The ten-dimensional chiral spinor $\Psi$ also decomposes into
the gaugino $\psi\equiv \frac{1-\Gamma_{5678}}2 \Psi$
and hypermultiplet fermion $\chi\equiv \frac{1+\Gamma_{5678}}2 \Psi$.
Our spinor and gamma matrix conventions are summarized in
Appendix \ref{sec:spinor-gamma}.
Real fields are hermitian matrices,
and the gauge covariant derivative is $D_\mu=\partial_\mu+iA_\mu$.
In terms of the coordinates $x^\mu=(x^i,\tau)$
($\mu=1,\ldots,4$, $i=1,2,3$),
the metric is simply $ds^2=d\tau^2+d x^i dx^i$.
We denote the radius of the Euclidean time circle by $R$.
The theory is defined by the physical action
\begin{equation}
S=S_\text{vec}+S_\text{hyp}\,,
\label{total-action}
\end{equation}
where the two terms describing the vector and hypermultiplets are given by
\begin{eqnarray}
S_\text{vec}
&=&\frac {1}{g^2} \int_{S^1\times \mathbb R^3}
d^4x\,
{\rm Tr} \left (
\frac 1 2
F_{\mu\nu} F^{\mu\nu}
+
D_\mu \Phi_A D^\mu \Phi_A
-[\Phi_0,\Phi_9]^2
-\psi \Gamma^\mu D_\mu\psi
- i \psi \Gamma^A [\Phi_A, \psi]
\right)
\nonumber\\
&&\quad
+\frac{i\vartheta}{8\pi^2} \int_{S^1\times \mathbb R^3}{\rm Tr} \left( F\wedge F\right)
\,,
\label{action-vec}
\end{eqnarray}
and
\begin{eqnarray}
S_\text{hyp}
&=&\frac {1}{g^2} \int_{S^1\times \mathbb R^3}
d^4x\,
{\rm Tr} \bigg(
D_\mu \Phi_i D^\mu \Phi_i
-\frac 1 2[\Phi_i,\Phi_j]^2
-\left(
[\Phi_A, \Phi_i]
-i M_{A\,ij}\Phi_j
\right)^2
-\chi \Gamma^\mu D_\mu\chi
\nonumber\\
&&\quad\quad\quad\quad\quad\quad\quad\quad\quad
- i \chi
\Gamma^A \left([\Phi_A, \chi] -\frac{i}{4} M_{A\,ij}\Gamma^{ij} \chi
\right)
- i \chi \Gamma^i [\Phi_i, \chi]
\bigg)
\,.
\label{action-hyp}
\end{eqnarray}
Here ${\rm Tr}$ denotes an invariant metric on the Lie algebra of
the gauge group $G$, $\vartheta$ is the theta angle,
and $i,j=5,6,7,8$ denote the hypermultiplet
scalar directions.
The two real anti-symmetric matrices $M_{ij}\equiv M_{0\,ij}$ and $M_{9\,ij}$ are
proportional to a single pure-imaginary anti-symmetric matrix
$F_{ij}$%
\footnote{%
The flavor symmetry generator $F_{ij}$ ($i,j=5,\ldots,7$)
should not be confused with the field strength
$F_{MN}=-i [D_M,D_N]$ ($M,N=1,\ldots,9,0$).
}, which is normalized
as $F_{ij}F_{ji}=4$ and is taken to be anti-self-dual in the $5678$ directions
so that only the hypermultiplet fermions get massive.
The flavor generator $F$ is represented as $F_{ij}$
on the scalars and as $\frac 1 4 F_{ij}\Gamma^{ij}$ on spinors.
The real mass parameters $M\equiv M_0$ and $M_9$ are defined by
$M_{A\,ij}=i M_A F_{ij}$ ($A=0,9$).
The massless limit is $\mathcal N=4$ super Yang-Mills.
Our aim is to compute the expectation value
of half-BPS line operators along $S^1$,
placed at a point on the 3-axis of $\mathbb R^3$.
The most basic line operator
is the Wilson operator defined as
\begin{eqnarray}
W_R={\rm Tr}_R P \exp\oint_{S^1}\left(-iA+\Phi_0\right) d\tau\,.
\label{wilson-def}
\end{eqnarray}
This is labeled by the representation $R$ of the gauge group, or equivalently
its highest weight.
The supersymmetric 't Hooft operator with charge $B$ is defined
by integrating over the fluctuations of the fields
around the configuration
\begin{equation}
\begin{aligned}
A&\equiv A_\mu dx^\mu
=\left(ig^2\vartheta\frac{B}{16\pi^2}\frac{1}{r} +{A_{\tau}^{(\infty)}} \right)d\tau
+\frac B 2 \cos\theta d\varphi \\
\Phi_0&=
- g^2\vartheta \frac{B}{16\pi^2}\frac{1}{r}+\Phi_0^{(\infty)}\,,
\quad\quad
\Phi_9= \frac B{2r}+\Phi_9^{(\infty)}
\end{aligned}
\quad\quad\text{ in the background.}
\label{thooft-background}
\end{equation}
We recall that $\tau\equiv x^4$ and that $\vartheta$ is the gauge theory theta angle. We have also introduced polar coordinates
$(r\equiv |\vec x|,\theta, \varphi)$ for $\mathbb R^3$.
Our choice of scalars in (\ref{wilson-def}) and (\ref{thooft-background})
ensures that the Wilson and 't Hooft operators
preserve the same sets of supercharges.
The action of the $U(1)$ R-symmetry
rotates $\Phi_0+i\Phi_9$ and changes
the set of preserved supercharges.
Note that we define the electric Wilson line $A_\tau^{(\infty)}$
in the local trivialization such that the $d\varphi$ term is
given by $(B/2)\cos\theta d\varphi$ rather than the more familiar
$-(B/2)(\pm 1-\cos\theta)d\varphi$.
Our choice guarantees that when $\lambda\neq 0$,
the holonomy at the spatial infinity with $\theta=\pi/2$
is $\exp(-2\pi i R A_\tau^{(\infty)})$.
This will play a role in Section \ref{sec:non-com}.
More general line operators are dyonic and carry both electric and magnetic
charges.
Such operators are defined by a path integral for a 't Hooft operator
with charge $B$, with the insertion of a Wilson operator
for the stabilizer of $B$ in $G$.
The dyonic charges are elements of the sum of coweight and weight lattices of $G$
\begin{equation}
\Lambda_{cw}\oplus \Lambda_w\,,
\end{equation}
and the charges related by a simultaneous action of the Weyl group
the two lattices are equivalent
\cite{Kapustin:2005py}.
Due to Dirac quantization, the magnetic charge must be a coweight
which has integer inner products with all the
weights in the matter representation.\footnote{In the theories whose gauge group is a product of $SU(2)$'s,
the electric and magnetic charges with these constraints and equivalence
relations match the homotopy classes of non-self-intersecting
curves on the corresponding Riemann surface \cite{Drukker:2009tz}.
}
Having defined the line operators whose vevs we wish to compute,
let us explain the parameters of the theory those vevs will depend on.
We are studying the theory in the Coulomb branch, so
the real scalars in the vector multiplet have the expectation values
\begin{equation}
\langle \Phi_A\rangle \equiv \Phi_A^{(\infty)}\in \mathfrak t
\quad\quad
A=0\,,9\,,
\end{equation}
which are the asymptotic values at $|\vec{x}|=\infty$.
Since we compactify the theory on $S^1$,
we also have the electric and magnetic Wilson lines.
The electric Wilson line is the asymptotic value of the $\tau$-component of the
gauge field
\begin{equation}
A_\tau^{(\infty)}\in \mathfrak t\,.
\end{equation}
Due to potential terms in the action (\ref{action-vec}),
$\Phi_A^{(\infty)}$ and $A_\tau^{(\infty)}$ can be
simultaneously diagonalized, i.e.,
they can take values in the Cartan subalgebra $\mathfrak t$.
We also need to consider the magnetic Wilson line.
In the IR theory this is the vev of the scalar dual to the gauge field
in three dimensions.
In the UV theory we define it as follows.
At a generic point of the Coulomb branch, the scalar vevs $\Phi_A^{(\infty)}$ classically breaks the gauge group $G$ to the maximal torus $T$.
The path integral includes infinitely many sectors
classified by the magnetic charges at infinity.
The general boundary condition is such that
asymptotically as $|\vec x|\rightarrow \infty$,
we allow $\Phi_A(\vec x)$ to take
any values that are gauge equivalent to $\Phi_A^{(\infty)}$, i.e.,
there is a map $g: S^2\rightarrow G$ such that
\begin{eqnarray}
\Phi_A(\vec x)\rightarrow g(\vec n)\cdot \Phi_A^{(\infty)}\cdot
g^\mo(\vec n) \quad \text{ as }\quad |\vec x| \rightarrow \infty
\end{eqnarray}
with $\vec n\equiv \vec x /|\vec x|\in S^2$.
Then the scalars $\Phi_A(\vec x)|_{|\vec x|=\infty}$ themselves
define a map from $S^2$ to the orbit
$\{g
(\Phi_0^{(\infty)}, \Phi_9^{(\infty)})
g^{-1}|
g\in G\}$,
which is diffeomorphic to $G/T$
because the stabilizer of a generic element of $\mathfrak t\times \mathfrak t$
is $T$.
We can demand that $g=1$ at the north pole of $S^2$,
so that $\Phi_A$ at $|\vec x|=\infty$ define
a homotopy class in $\pi_2(G/T)$ with a base point at the north pole.
If $G$ is simply connected,
the maximal torus can be identified
with the quotient of the Cartan subalgebra
by the coroot lattice\footnote{See \cite{Gukov:2006jk} for a review of lattices in the Cartan subalgebra
$\mathfrak t$ and
its dual $\mathfrak t^*$.
}
$T\simeq \mathfrak{t}/\Lambda_{cr}$,
so $\pi_2(G/T)\simeq \pi_1(T) =\Lambda_{cr}$.
In fact $G/T$ depends only on the Lie algebra of $G$,
so $\pi_2(G/T)= \Lambda_{cr}$ for any $G$.
The infinitely many topological sectors are therefore classified by
$ \Lambda_{cr}$.
Physically this makes sense because $ \Lambda_{cr}$ is the lattice of magnetic charges carried by Polyakov-'t Hooft monopoles.
This lattice is more coarse than the coweight lattice
$\Lambda_{cw}$ in which the magnetic charge $B$ of the 't Hooft operator
takes values,
$\Lambda_{cr}\subset \Lambda_{cw}$.
With generic matter representations, the lattice of
't Hooft charges $B$ allowed by Dirac quantization
would be smaller than $\Lambda_{cw}$.
Let us now insert a 't Hooft operator with magnetic charge $B\in \Lambda_{cw}$
at the origin. The insertion of the 't Hooft operator
changes the topology of the vector bundles in which the
fields take values, and in particular
the structure of the boundary conditions at spatial infinity.
One can classify the allowed configurations by the asymptotic
magnetic charges taking values in the shifted lattice
$\Lambda_{cr}+B\subset \Lambda_{cw}$.
We define the magnetic Wilson line $\Theta\in \mathfrak t^*$ as
the chemical potential for the magnetic charges.
The expectation value of the 't Hooft operator
is given by the sum
\begin{eqnarray}
\langle T_B\rangle= \sum_{v\in \Lambda_{cr}+B}
e^{i v\cdot \Theta}
\int_v \mathcal D A
\mathcal D\Psi
\, e^{-S}\,,
\label{vev-sum-v}
\end{eqnarray}
where the path integral in each summand is performed with the boundary condition specified by $v$.
In the three-dimensional Abelian gauge theory that arises
via dimensional reduction, $\Theta$
is identified with
the expectation values of scalars dual to the photons
\cite{Affleck:1982as},
and the UV and IR definitions of $\Theta$ are consistent.
Along the circle $S^1$ we can impose various
twisted boundary conditions on the fields.
It is convenient to exhibit them by representing the line operator vev as a supersymmetric index, taking $S^1$ as a time direction.
The line operator $L$
modifies the Hilbert space of the theory, rather than
acts on the original Hilbert space as a linear transformation.
We define our observable,
the expectation value of the line operator $L$,
to be a trace in the modified Hilbert space $\mathcal H_L$
\begin{eqnarray}
\langle L\rangle= {\rm Tr}_{\mathcal H_L}
(-1)^{F}e^{-2\pi R H}e^{2\pi i\lambda(J_3+I_3)}
e^{2\pi i \mu_f F_f}\,,
\label{trace-def}
\end{eqnarray}
where $J_3$ and $I_3$ are the generators of the Lorentz $SU(2)$
and the R-symmetry $SU(2)$.
Here $J_3$ generates a rotation along the 3-axis:
$i J_3=
x^1 \partial_2-x^2 \partial_1$
when acting on a scalar.
As we will see below, the combination $J_3+I_3$ commutes with the supercharge we use for localization.
We have also included the twist by the flavor symmetries
with generators $F_f$
and chemical potentials $\mu_f$, $f=1,\ldots, N_\text{F}$.
The definition (\ref{trace-def}) of the line operator vev
coincides with the one used in \cite{Gaiotto:2010be}.
The system may be realized in terms of a path integral
over the fields with appropriate twisted boundary conditions along $S^1$.
In this paper we adopt the equivalent formulation
where everywhere in the action (\ref{total-action}) on $\mathbb R^4$ the time derivative
is shifted as
\begin{equation}
\partial_\tau \rightarrow
\partial_\tau
- \frac{i}{R} \lambda(J_3+I_3)
- \frac{i}{R}\sum_{f=1}^{N_\text{F}} \mu_f F_f
\label{der-shift}
\end{equation}
and the fields are periodic in $\tau$.
The electric and magnetic Wilson lines can also be regarded as the chemical potentials for the corresponding charges.
As we will see all the parameters except $\lambda$ will enter the line operator vevs in specific complex combinations.
These are the moduli
\begin{equation}
a\equiv R\, (A_\tau^{(\infty)}+i\Phi_0^{(\infty)})
\in \mathfrak t_{\mathbb C}\,,
\quad
b\equiv
\frac{\Theta}{2\pi }- \frac{4\pi i R}{g^2}\Phi_9^{(\infty)}
+\frac{\vartheta}{2\pi} a
\in \mathfrak t^*_{\mathbb C}
\,.\label{def-ab}
\end{equation}
and the complexified mass parameters
\begin{equation}
m_f\equiv -\mu_f+iR M_f\in \mathbb C
\quad\quad
f=1,\ldots, N_\text{F}\,.
\end{equation}
We use the Lie algebra metric ${\rm Tr}$
in the action to regard $\Phi_9^{(\infty)}$
and $a$
as elements of $\mathfrak t^*_{\mathbb C}$.
General $\mathcal N=2$ theories
have several mass parameters $M_{Af}$ with
$A=0,9$ and $f=1,\ldots, N_\text{F}$.
These can be thought of as the vevs of the scalars
in the vector multiplets
that weakly gauge the flavor symmetries.
Only $M_f\equiv M_{A=0,f}$, which are the analog of $\Phi_0$,
will enter the line operator vevs.
\section{Localization for gauge theories on $S^1\times \mathbb R^3$}
\label{sec:localization}
We apply the localization technique
introduced for calculations in gauge theory on $S^4$ \cite{Pestun:2007rz}.
In this formalism, one adds a new term $t Q\cdot V$ to the action,
so that the path integral takes the form
\begin{eqnarray}
\int {\mathcal D} A{\mathcal D}\Psi e^{-S-t Q\cdot V}\,.
\end{eqnarray}
Here $A$ and $\Psi$ include all the bosons and fermions, respectively.
We will also need to add ghost fields after gauge-fixing.
For observables that are invariant under the supercharge $Q$ of choice,
the path integral is independent of the parameter $t$.
The localization action is chosen to be
$V=(\Psi,\overline{Q\cdot \Psi})=(\psi,\overline{Q\cdot\psi})
+(\chi,\overline{Q\cdot\chi})$, where $\psi$ and $\chi$ denote
the fermions in the vector multiplet and the hypermultiplet.
Since the bosonic part of $Q\cdot V$ is a positive definite term
$||Q\cdot \Psi||^2$, the path integral is dominated by the solutions
of $Q\cdot \Psi=0$ in the limit $t\rightarrow +\infty$ and can be calculated exactly by summing the fluctuation determinants at all the saddle points.
\subsection{Symmetries}
For localization we need to close off-shell the relevant subalgebra
of the whole superalgebra.
For this we introduce seven auxiliary fields $K_j$ as in \cite{Pestun:2007rz}.
The supersymmetry transformations in $\mathcal N=2^*$ theory are given by
\begin{eqnarray}
Q\cdot A_M&=& \epsilon \Gamma_M \Psi\,,
\\
Q\cdot\Psi&=&\frac 1 2 F_{MN} \Gamma^{MN} \epsilon+i K^i\nu_i\,,
\\
Q\cdot K_j &=& i\nu_j \Gamma^M D_M \Psi\,.
\end{eqnarray}
The gamma matrices and the constant spinors $\nu_i$ ($i=1,\ldots,7$) are defined in Appendix \ref{sec:spinor-gamma}.
The gauge fields in $F_{MN}$ and $D_M$ include mass matrices
$M_{Aij}=i M_A F_{ij}$ through the Scherk-Schwarz mechanism \cite{Pestun:2007rz}.
The spinor $\epsilon$ must be chosen so that the line operators are invariant under the supersymmetry transformation $Q$. We will use the same spinor as used in
\cite{Gomis:2011pf}
\begin{equation}
\epsilon
= \frac{1}{\sqrt 2}(1,0^7,1,0^7) \,,
\label{spinor-chosen}
\end{equation}
where the power indicates the number of repeated entries.
It satisfies\footnote{\label{foot:Nek}%
The third condition implies that $Q$ corresponds to the fermionic symmetry for the Donaldson-Witten twist \cite{Witten:1988ze} in the 1239-directions. Thus $\langle L\rangle$ is a limit of the five-dimensional Nekrasov partition function \cite{Nekrasov:2002qd} for a theory on $S^1\times \mathbb R^4$ with a line operator insertion, where one of the equivariant parameter for the rotation in the 39 plane
is set to zero and a direction in $\mathbb R^4$ is compactified on an
infinitely small circle.
}
\begin{equation}
\Gamma_{5678}\epsilon=-\epsilon\,,
\quad
\Gamma_{04}\epsilon=-i\epsilon\,,
\quad
\Gamma_{1239}\epsilon=\epsilon\,,
\quad
(2\Gamma_{12}+\Gamma_{56}+\Gamma_{78})\epsilon=0\,.
\end{equation}
The last condition implies that the supercharge commutes with the combination $J_3+I_3$ of spatial and R-symmetry rotations.
This explains why this particular combination entered the definition (\ref{trace-def}) of the vev.
We will need later the square of the supersymmetry transformation
given by
the spinor $\epsilon$ in (\ref{spinor-chosen}),
Using the vector
\begin{eqnarray}
v^M\equiv \epsilon \Gamma^M \epsilon=(i,0^3,1,0^5)\quad\quad
M=0,1,\ldots,9\,,
\end{eqnarray}
we find that $Q^2$ generates time translation,
minus the complexified gauge transformation $G_\Lambda$
with gauge parameter $\Lambda=A_\tau+i\Phi_0$,
and the flavor symmetry transformation $iMF$:
\begin{eqnarray}
Q^2 \cdot A_M&=&- F_{\tau M}-[i \Phi_0, D_M] -i \delta_M^i M_{ij}\Phi_{j}
\,,\nonumber\\
Q^2\cdot \Psi&=&- \partial_\tau \Psi-i [ A_\tau+i\Phi_0,\Psi]
-\frac i 4 M_{ij} \Gamma^{ij}\Psi
\,,\label{Q2APsiK}\\
Q^2\cdot K_i&=&- \partial_\tau K^i-i[A_\tau+i\Phi_0,K_i]
\,.\nonumber
\end{eqnarray}
See Appendix C and (2.27) of \cite{Pestun:2007rz}.
\subsection{Localization equations}
\label{sec:loc-eq}
Let us study the localization equations $Q\cdot \Psi=0$,
whose solutions the path integral localizes to.
We decompose $\Psi$ as
\begin{eqnarray}
\Psi=\sum_{M=1}^9\Psi_M\tilde \Gamma^M\bar \epsilon
+i\sum_{j=1}^7 \Upsilon_j \nu^j\,.
\end{eqnarray}
Noting that
\begin{eqnarray}
\Psi_M=\epsilon\Gamma_M \Psi\,,~~~~~
i \Upsilon_j=\bar \nu_j \Psi\,.
\end{eqnarray}
we obtain
\begin{eqnarray}
0&=& Q\cdot\Psi_M= \frac 1 2 F_{PQ}\,\epsilon \Gamma_M \Gamma^{PQ} \epsilon
\quad\quad
M=1,\ldots,9\,,
\label{QPsiM}
\\
0&=& i Q\cdot \Upsilon_j= \frac 12 F_{MN}\, \bar\nu_j \Gamma^{MN}\epsilon
+i K_j
\quad\quad\quad
j=1,\ldots,7\,.
\label{QUpsilon}
\end{eqnarray}
The equations (\ref{QPsiM}) reduce to\footnote{To show this we used the identities $\Gamma_{M} \tilde \Gamma_{[P} \Gamma_{Q]}=\Gamma_{[M} \tilde \Gamma_P \Gamma_{Q]} +2 \delta_{M[P} \Gamma_{Q]}$ and $\epsilon \Gamma_{[M} \tilde \Gamma_P \Gamma_{Q]}\epsilon=0$.
}
\begin{eqnarray}
0= Q\cdot \Psi_M= -v^N F_{NM}\,.
\end{eqnarray}
According to (\ref{Q2APsiK}), these are equivalent to $Q^2$-invariance,
i.e., invariance under a combination of $\tau$-translation,
gauge transformations, and flavor transformations.
Due to the replacement of the $\tau$-derivative in (\ref{der-shift}),
for generic $\lambda$ the bosonic fields must also be invariant under the combination $J_3+I_3$ of spatial and R-symmetry rotations.
Among the various components of (\ref{QUpsilon}),
the most important equations are\footnote{We used the following facts:
$\bar \nu_j \Gamma^{kl}\epsilon=-\epsilon_{jkl}$ for $j,k,l\in\{1,2,3\}$,
$\bar\nu_j \Gamma^{kl}\epsilon=0$ for $j,k\in\{1,2,3\}$ and $l\in\{5,6,7,8\}$,
$\bar\nu_j \Gamma^{k9}\epsilon=\delta_{jk}$ for $j,k\in\{1,2,3\}$, and
$\bar\nu_j \Gamma^{9l}\epsilon=0$ for $j\in\{1,2,3\}$ and $l\in\{5,6,7,8\}$.
We also went ahead and set the hypermultiplets to zero.
This is justified below by $Q^2$-invariance.
}
\begin{equation}
0=i Q\cdot \Upsilon_j= D_j \Phi_9- \frac{1}{2}\sum_{k,l=1}^3 \epsilon_{jkl}F_{kl}
+i K_j
\quad\quad
j,k,l=1,2,3\,.
\end{equation}
The imaginary part sets $K_j$ to zero.
The real part is precisely the Bogomolny equations
\begin{equation}
*_3 F=D \Phi_9
\label{eq:Bogo}
\end{equation}
that describe monopoles on $\mathbb R^3$! Thus we conclude that the path integral localizes to the fixed points on the monopole moduli space with respect to spatial rotations and gauge transformations.
Four other components of (\ref{QUpsilon}) read\begin{equation}
\begin{aligned}
0=
i Q\cdot \Upsilon_j= \sum_{k=1}^3 \sum_{l=5}^8(\bar\nu_j \Gamma^{kl} \epsilon) D_k \Phi_l
+ \sum_{l=5}^8 (\bar\nu_j \Gamma^{9l} \epsilon)i[\Phi_9, \Phi_{l}]
+i K_j
\quad\quad
j=4,5,6,7\,.
\end{aligned}
\label{QPsij4-7}
\end{equation}
Again the imaginary part requires $K_j$ to vanish.
The real part of (\ref{QPsij4-7}) is in fact the ``realification''
of the Dirac-Higgs equation
\begin{equation}
\sum_{i=1}^3
\sigma^i D_{i} q +
[\Phi_9, q]=0\,,
\label{DH-eq}
\end{equation}
where the two-component ``spinor'' $q$ is a linear combination of $\Phi_i$
with $i=5,6,7,8$.
See Appendix \ref{sec:diff-op} for a related discussion.
As in topological twist, the hypermultiplet scalars behave as a spinor under the combination $J_3+I_3$.
Though generically (\ref{DH-eq}) itself admits non-zero solutions,
the $Q^2$-invariance, in particular the invariance under flavor transformations, requires $q$ to vanish.
Thus localization on $S^1\times \mathbb R^3$ leaves no bosonic zero-mode to be integrated over, and the final answer for the vev will
be expressed as a finite sum. This is in contrast with the results for $S^4$ \cite{Pestun:2007rz,Gomis:2011pf} where the path integral reduced to a finite dimensional matrix integral.
\subsection{On-shell action}
\label{sec:classical}
Let us work out the classical contribution $e^{-S_\text{cl}}$,
given by the on-shell action evaluated in the background
(\ref{thooft-background}).
The on-shell action for the hypermultiplet simply vanishes,
therefore we focus on the action (\ref{action-vec}) for the vector multiplet.
For the background (\ref{thooft-background}), we also have
\begin{eqnarray}
F&=&ig^2\vartheta \frac{B}{16\pi^2 } \frac{d\tau\wedge dr}{r^2}-\frac B 2
\sin\theta d\theta\wedge d\varphi
\,,\\
*F&=&-\frac B {2r^2} d\tau\wedge dr +ig^2\vartheta \frac{B}{16\pi^2 }
\sin\theta d\theta\wedge d\varphi
\,.
\end{eqnarray}
Our orientation is such that the volume form is
$d\tau\wedge dx^1\wedge dx^2\wedge dx^3$.
The action (\ref{action-vec}) is divergent in the presence of
such a singular dyonic background.
We can render the action finite
by cutting off the spacetime at $\Sigma_3\equiv\{r=\delta\}$ and
by adding the boundary term \cite{Giombi:2009ek,Gomis:2011pf}
\begin{eqnarray}
S_\text{bdry}
&=&
\frac 2{g^2}\int_{\Sigma_3} {\rm Tr} \left( \Phi_9 F -i \Phi_0 * F\right)
\wedge d\tau
\,.
\label{boundary-action}
\end{eqnarray}
We find that
\begin{equation}
\begin{aligned}
S_\text{vec}=&
\frac{1}{g^2 \delta}\left(4\pi^2 R+\frac{g^2\vartheta^2 R}{16\pi^2}\right)
{\rm Tr} B^2-i\vartheta R\,{\rm Tr}\left(A_\tau^{(\infty)}B\right)\,,\\
S_\text{bdry}=&
-\frac{1}{g^2 \delta}\left(4\pi^2 R+\frac{g^2\vartheta^2 R}{16\pi^2}\right)
{\rm Tr} B^2
-\frac{8\pi^2 R}{g^2}{\rm Tr}\left(\Phi_9^{(\infty)} B\right)
+\vartheta R {\rm Tr}\left( \Phi_0^{(\infty)} B\right)\,.
\end{aligned}
\end{equation}
Thus the classical on-shell action is given by
\begin{eqnarray}
S_\text{cl}(B)\equiv S_\text{vec}+S_\text{bdry}
&=&
-\frac{8\pi^2 R }{g^2}
{\rm Tr}
\left[
\Phi_9^{(\infty)}
B
\right]
-
i\vartheta R\ {\rm Tr} \left[ \left( A_\tau^{(\infty)}+
i\Phi_0^{(\infty)}\right)
B \right]\,.
\label{on-shell-action-result}
\end{eqnarray}
The on-shell action nicely combines with the weight $e^{i B\cdot \Theta}$
for the magnetic charge in (\ref{vev-sum-v}) so that
\begin{equation}
\langle T_B\rangle\sim e^{i B\cdot\Theta} e^{-S_\text{cl}(B)}
=e^{2\pi i B\cdot b}\,,
\end{equation}
where $b$ was define in (\ref{def-ab}).
This is the leading classical approximation to the 't Hooft operator vev.
We will compute one-loop and non-perturbative corrections in the following sections.
\section{One-loop determinants}
\label{sec:one-loop}
Having computed the classical contribution to the 't Hooft operator vev, in this section we will compute the one-loop correction
following \cite{Pestun:2007rz} and in parallel with \cite{Gomis:2011pf}. As we saw in the previous section, the path integral reduces to a sum over saddle points. For each saddle point we need to compute the fluctuation determinants. The methods here will also be used in Section \ref{sec:mono} for the computation of such non-perturbative corrections.
\subsection{Gauge fixing}
The gauge fixing action in the $R_\xi$-gauge is
\begin{eqnarray}
S_\text{gf}=\int d^4x\,
{\rm Tr}
\left (
-i\,
\tilde c\,
\sum_{M=1,2,3,9}D_{(0)}^M D_M c
+ \tilde b\left(i
\sum_{M=1,2,3,9}D_{(0)}^M
\tilde A_M+\frac \xi 2 \tilde b\right)
\right)\,.
\end{eqnarray}
We have defined $\tilde A_M\equiv A_M-A_{(0)M}$ where
$A_{(0)M}$ is the background configuration given in (\ref{thooft-background}).
The ghost fields $c, \tilde c$ are fermionic, and $\tilde b$ is bosonic.
By defining the BRST transformations\footnote{To compare with Pestun's formalism in \cite{Pestun:2007rz},
set $\tilde a_0, b_0,c_0, \tilde c_0$ to zero.
Then separate his BRST transformation $\delta$ into our $Q_\text{B}$
and the part $\delta_0$ proportional to $a_0$: $\delta= Q_\text{B}+\delta_0$.
Then our $Q$ can be written as $s+\delta_0$ with $a_0=-\Phi_{(0)0}$,
where $s$ denotes the supersymmetry transformation in \cite{Pestun:2007rz}.
}
\begin{equation}
\begin{aligned}
& Q_\text{B}\cdot A_M=- [c, D_M]\,,\quad
Q_\text{B}\cdot \Psi=-i [c, \Psi]\,,\quad
Q_\text{B}\cdot K_i=- i[c, K_i]\,,\quad
\\
&
Q_\text{B}\cdot c=-\frac i 2 [c,c]\,,\quad
Q_\text{B} \cdot\tilde c=\tilde b\,,\quad
Q_\text{B}\cdot \tilde b=0\,,
\end{aligned}\end{equation}
we can write
\begin{eqnarray}
S_\text{gf}=
Q_\text{B}\cdot V_\text{gh}\,,
\quad
V_\text{gh}\equiv
\int d^4x
{\rm Tr}
\left(
\tilde c\left(i\,
\sum_{M=1,2,3,9}D_{(0)}^M \tilde A_M
+\frac \xi 2 \tilde b\right)
\right)\,.
\end{eqnarray}
The BRST transformation squares to zero, $\{Q_\text{B},Q_\text{B}\}=0$.
Unlike the case of $S^4$ \cite{Pestun:2007rz}
where the spacetime is compact,
we do not need to introduce ghosts-for-ghosts to deal with
constant gauge transformations.
We define the action of the supercharge $Q$ on the ghosts by
\begin{equation}
\begin{aligned}
& Q\cdot c=- v^M \tilde A_M\equiv -\tilde \Phi=-i \tilde \Phi_0- \tilde A_\tau\,,\quad
Q\cdot \tilde c=0\,,
\\
&
Q\cdot \tilde b=- v^M D_M\tilde c=-\partial_\tau \tilde c
-i [A_\tau+i\Phi_0,\tilde c]\,.
\end{aligned}
\end{equation}
In the background $Q$ annihilates all the fermions, therefore the background
is supersymmetric.
We have $\{Q,Q\}(\text{ghost})=0$.
\subsection{One-loop determinants and the index theorem}
After gauge fixing, the total fermionic symmetry we use for localization is
\begin{equation}
\hat Q\equiv Q+Q_\text{B}\,.
\end{equation}
While $Q^2$ in (\ref{Q2APsiK}) involves a gauge transformation $G_\Lambda$
with a dynamical gauge parameter $\Lambda=A_\tau+i\Phi_0$,
the gauge transformation that appears in $\hat Q^2=Q^2+\{Q,Q_\text{B}\}$
turns out to have a fixed parameter
$\Lambda=A_{(0)\tau}+i\Phi_{(0)0}=A^{(\infty)}_\tau+i\Phi^{(\infty)}_0$:\footnote{For the gauge field
$\hat Q^2\cdot A_M=
-\partial_\tau \tilde A_M -i[A^{(\infty)}_\tau+i\Phi^{(\infty)}_0, \tilde A_M]$.
}
\begin{eqnarray}
\hat Q^2=-
\partial_\tau
-i (A_\tau^{(\infty)}+i\Phi_0^{(\infty)})
+M F\,.
\label{Qhat-squared}
\end{eqnarray}
Saddle points of the path integral remain the same
after we replace $Q\cdot V$ by $\hat Q\cdot \hat V$.
Recall that $M\equiv M_0$ is one of the mass parameters
defined below (\ref{action-hyp})
and that $F$ is the flavor symmetry generator.
The path integral to consider is
\begin{eqnarray}
\int \mathcal D A \mathcal D\Psi \mathcal DK
\mathcal D \tilde b
\mathcal D c
\mathcal D \tilde c\,
e^{-S-t \hat Q\cdot \hat V}\,,
\end{eqnarray}
where
\begin{equation}
\hat V
=
\left\langle \Psi\,,\,
\overline{\hat Q\cdot \Psi}
\right\rangle
+V_\text{gh}\,.
\end{equation}
In order to evaluate the path integral in the limit $t\rightarrow \infty$, we need to compute the superdeterminant of the kinetic operator in $\hat Q_{(0)}\cdot \hat V^{(2)}$, where $\hat Q_{(0)}$ is the linearization of $\hat Q$, and $ \hat V^{(2)}$ is the quadratic part of $\hat V$.
Following \cite{Pestun:2007rz} let us define
\begin{eqnarray}
X_0=(\tilde A_M)_{M=1}^9\,,\quad
X_1=(\Upsilon_i, c, \tilde c)
\label{X-def}
\end{eqnarray}
and their partners
\begin{equation}
\begin{aligned}
X_0'\equiv& \hat Q_{(0)}\cdot X_0=(\Psi_M-[c, D_{(0)M}])_{M=1}^9\,,
\\
X_1'\equiv& \hat Q_{(0)}\cdot X_1=
\left(
K_i-i(\bar\nu_i\Gamma^{MN}\epsilon)
D_{(0)M} \tilde A_N
\,
, -\tilde\Phi,b\right)\,.
\end{aligned}
\label{Xprime-def}
\end{equation}
Now $\hat V^{(2)}$ takes the form
\begin{eqnarray}
V^{(2)}
=
\left\langle
\left(
\begin{array}{cc}
X'_0& X_1
\end{array}
\right)
\,,
\left(
\begin{array}{cc}
D_{00}&D_{01}\\
D_{10}&D_{11}
\end{array}
\right)
\left(
\begin{array}{c}
X_0\\
X'_1
\end{array}
\right)
\right\rangle
\,,
\label{eq:V2}
\end{eqnarray}
where $D_{00}$ and others are certain differential operators.
Then ${\hat Q}_{(0)}\cdot V^{(2)}$ is given by
\begin{equation}
\begin{aligned}
\hat{Q}_{(0)}\cdot V^{(2)}=&
\left\langle
\begin{pmatrix}
X_0,X_1'
\end{pmatrix}
\begin{pmatrix}
-\mathcal R_{00}&\\
&1
\end{pmatrix}
\,,
\begin{pmatrix}
D_{00}&D_{01}\\
D_{10}&D_{11}
\end{pmatrix}
\begin{pmatrix}
X_0\\
X_1'
\end{pmatrix}
\right\rangle
\\
&
+
\left\langle
\begin{pmatrix}
X_0',X_1
\end{pmatrix}
\,,
\begin{pmatrix}
D_{00}&D_{01}\\
D_{10}&D_{11}
\end{pmatrix}
\begin{pmatrix}
-1&\\
&-\mathcal R_{11}
\end{pmatrix}
\begin{pmatrix}
X_0'\\
X_1
\end{pmatrix}
\right\rangle
\,,
\end{aligned}
\end{equation}
where $\hat Q_{(0)}^2\cdot X_0=\mathcal R_{00}\cdot X_0$
and
$\hat Q_{(0)}^2\cdot X_1=\mathcal R_{11}\cdot X_1$.
Thus the one-loop determinant is given by
\begin{eqnarray}
Z_\text{1-loop}&=&
\frac{\det^{1/2}\left[
\begin{pmatrix}
D_{00}&D_{01}\\
D_{10}&D_{11}
\end{pmatrix}
\begin{pmatrix}
- 1&\\
&-\mathcal R_{11}
\end{pmatrix}
\right]}{\det^{1/2}
\left[
\begin{pmatrix}
-\mathcal R_{00}&\\
&1
\end{pmatrix}
\begin{pmatrix}
D_{00}&D_{01}\\
D_{10}&D_{11}
\end{pmatrix}
\right]
}
=
\frac{\det^{1/2} \mathcal R_{11}}
{\det^{1/2}\mathcal R_{00}}
\nonumber\\
&=&
\frac{\det^{1/2}_{\text{Coker} D_{10}} \mathcal R}
{\det^{1/2}_{\text{Ker} D_{10}}\mathcal R}
\,.
\label{Z-detR}
\end{eqnarray}
In the final line we have introduced notation $\mathcal R=\hat Q^2_{(0)}$ and used the fact that $\mathcal R$ commutes with $D_{10}$ as guaranteed by $\mathcal R$-invariance of $\hat V$.
Thus we only need the differential operator $D_{10}$, which can be obtained by explicitly computing $\hat V^{(2)}$.
It is easy to see what to expect from the results in Section \ref{sec:loc-eq}.
There we saw that the localization equations are given by the Bogomolny and Dirac-Higgs equations. In Appendix \ref{sec:diff-op}, we will show that $D_{10}$ involves the linearization of these equations as well as the dual of the gauge transformation.
The symmetry generator $\mathcal R=\hat Q^2_{(0)}$ is given in
(\ref{Qhat-squared}).
In a general $\mathcal N=2$ theory, we replace that last term $M F$ by $\sum_f M_f F_f$, where $F_f$ are the flavor symmetry generators in (\ref{trace-def}).
We also perform the shift (\ref{der-shift}) of the $\tau$ derivative.
It is also useful to rescale $\mathcal R$ as $\mathcal R\rightarrow -R\, \mathcal R$.
This does not affect the value of the one-loop determinant
(\ref{Z-detR}) due to cancellations between the numerator and the denominator.
Then $\mathcal R$ takes a simple expression
\begin{equation}
\mathcal R
= \varepsilon R\partial_\tau -i \lambda(J_3+I_3) + i a +i \sum_{f=1}^{N_\text{F}} m_f F_f\,.
\end{equation}
We have introduced a formal parameter $\varepsilon$
that should be set to one at the end of calculation.
A Fourier mode $e^{in\tau/R}$ along $S^1$ contributes $i n\varepsilon$
to $\mathcal R$.
The form (\ref{Z-detR}) of the one-loop determinant implies that
it can be obtained from the equivariant index of the operator $D_{10}$
\begin{equation}
\text{ind}\, D_{10}\equiv
{\rm Tr}_{\text{Ker} D_{10}} e^{2\pi \mathcal R}
-
{\rm Tr}_{\text{Coker} D_{10}} e^{2\pi \mathcal R}\,.
\end{equation}
Indeed if it is given in terms of weights $w_j$ and multiplicities
$c_j$ as
\begin{equation}
\text{ind}\, D_{10}=\sum c_j e^{w_j}\,,
\end{equation}
the one-loop determinant is given by
$
Z_\text{1-loop}=\left(\prod_j w_j^{c_j}\right)^{-1/2}.
$
In the following we will separately define the indices for differential operators acting on vector and hypermultiplets. We will also adopt a normalization
for $\text{ind}$
that corresponds to $ \text{ind}( D_{10})\rightarrow - \frac 1 2 \text{ind}( D_{10})$,
so that the translation from the index to the one-loop determinant
is simply given by the rule $\sum_j c_j e^{w_j}
\rightarrow \prod_j w_j^{c_j}$.
Then
\begin{equation}
Z_\text{1-loop}=\prod_j w_j^{c_j}\,.
\end{equation}
Thus we need to compute the weights under
the gauge transformation with parameter
$a\equiv R( A_\tau^{(\infty)}+i\Phi_0^{(\infty)})$,
a time translation by $\varepsilon$, and
a spatial rotation along the 3-axis with angle $2\pi\lambda$,
and flavor transformations with parameters $m_f$.
\subsection{Calculation of the equivalent index}
\label{sec:calc-index}
Before we delve into the details of the calculations, let us
summarize our methodology
that extends the techniques developed in \cite{Gomis:2011pf},
listing at the same time
the relevant complexes and their interrelations.
We showed above that the vector multiplet contribution to the one-loop determinant
can be computed from the index of the
complex that linearizes the Bogomolny equations in $\mathbb R^3$
\begin{equation}
D_\text{Bogo}: 0 \rightarrow \Omega^0(\text{ad}\,E)
\stackrel{
\text{\raisebox{-1mm}[0mm][0mm]{$(D\,, [i\Phi_9,\,\bullet\,])$}}
}{
\xrightarrow{\hspace*{1.6cm}}}
\Omega^1(\text{ad}\,E)\oplus \Omega^0(\text{ad}\,E)
\rightarrow\Omega^1(\text{ad}\,E)\rightarrow 0\,,
\label{complex-Bogo}
\end{equation}
where $\text{ad}\,E$ is the adjoint gauge bundle.
The second arrow is the gauge transformation whose conjugate\footnote{The equivariant index remains the same when we ``fold''
(\ref{complex-Bogo}) into
$ 0\rightarrow \Omega^0 \oplus \Omega^1 \rightarrow
\Omega^1\oplus \Omega^0\rightarrow 0$,
where twisting by $\text{ad}\,E$ is implicit, and
the second arrow is the linearized Bogomolny equations
plus the dual of a gauge transformation (\ref{dual-Bogo}).
The same remark applies to the self-dual complex (\ref{complex-SD}).
It is the folded form of the complexes that naturally arises from gauge-fixing.
}
appeared in (\ref{dual-Bogo}),
and the third is the map
$(\delta A,\delta\Phi_9)
\mapsto *D\delta A-D\delta\Phi_9+i[\Phi_9, \delta A]$
in (\ref{DBogo}).
As reviewed in Appendix \ref{sec:mono-inst},
the Bogomolny equations with a single
singularity on $\mathbb R^3$ are equivalent to the
anti-self-duality equations on the (single-centered) Taub-NUT space
with invariance under the action of
the group that we call $U(1)_K$.
Linearizing the correspondence,
we will obtain the index of the Bogomolny complex\footnote{We will refer to
(\ref{complex-Bogo}) and
(\ref{complex-DH})
as the Bogomolny and Dirac-Higgs (DH) complexes.
} (\ref{complex-Bogo})
from the index of the self-dual complex
\begin{equation}
D_\text{SD}:
0\rightarrow
\Omega^0
(\text{ad}\,E)
\stackrel{D}{\rightarrow}
\Omega^1(\text{ad}\,E)
\stackrel{
\text{\raisebox{-1mm}[0mm][0mm]{$(1+*) D$}}
}{
\xrightarrow{\hspace*{12mm}}}
\Omega^{2+}(\text{ad}\,E)
\rightarrow
0
\label{complex-SD}
\end{equation}
on the four-dimensional space
by taking an invariant part under the
$U(1)_K$ action \cite{MR1624279,Gomis:2011pf}.
Similarly the hypermultiplet contribution
will be derived from index of
the complex
\begin{equation}
D_{\text{DH},R}: 0\rightarrow
\Gamma(S\otimes R(E))
\stackrel{\text{\raisebox{-1mm}[0mm][0mm]{$\sigma^jD_j+\Phi_9$}}}{
\xrightarrow{\hspace*{13mm}}
}
\Gamma(S\otimes R(E))
\rightarrow 0\,,
\label{complex-DH}
\end{equation}
where $S$ is the spinor bundle over $\mathbb R^3$,
and $\Phi_9$ acts on $q\in \Gamma(S\otimes R(E))$ in the matter representation $R$.
Its index will be obtained from the $U(1)_K$ invariant
part of the index of the twisted Dirac complex
\cite{Gomis:2011pf}
\begin{equation}
D_{\text{Dirac},R}: 0\rightarrow \Gamma(S^+\otimes R(E)) \stackrel{
\text{\raisebox{-1mm}[0mm][0mm]{$ \bar \sigma^\mu D_\mu$}}
}{
\xrightarrow{\hspace*{9mm}}
} \Gamma(S^- \otimes R(E)) \rightarrow 0
\label{complex-Dirac}
\end{equation}
in four dimensions.
Both the self-dual and Dirac complexes are related to the Dolbeault complex
\begin{equation}
\bar D_R: 0\rightarrow \Omega^{0,0}(R(E))\rightarrow \Omega^{0,1}(R(E))
\rightarrow \Omega^{0,2}(R(E))
\rightarrow 0\,.
\label{complex-Dolb}
\end{equation}
To see this note that upon complexification we have
$\Omega^0_\mathbb{C}=\Omega^{0,0}$,
$\Omega^1_\mathbb{C}=\Omega^{1,0}\oplus \Omega^{0,1}$
and
$\Omega^{2+}_{ \mathbb C}=\Omega^{2,0}\oplus \Omega^{0,0}\omega
\oplus \Omega^{0,2}$, where $\omega$ is the K\"{a}hler form.
See, e.g., \cite{MR1079726}.
Since by Hodge duality
$\Omega^{2,2}=\Omega^{0,0}$ and $\Omega^{2,1}=\Omega^{1,0}$,
the complexification of the self-dual complex (\ref{complex-SD}) is isomorphic
to the Dolbeault complex (\ref{complex-Dolb}) with $R=\text{ad}$
twisted by $\Omega^{0,0}\oplus \Omega^{2,0}$.
For spinors recall that $\Omega^{p,q}=\Gamma(\Lambda^{p,q})$
and that $K=\Lambda^{2,0}$ is the canonical line bundle.
We have
\begin{equation}
S^+=K^{1/2}\otimes( \Lambda^{0,0}\oplus \Lambda^{0,2})\,,
\quad
S^-=K^{1/2}\otimes \Lambda^{0,1}\,.
\end{equation}
Thus the Dirac complex (\ref{complex-Dirac})
is isomorphic to the Dolbeault complex (\ref{complex-Dolb})
twisted by $(\Omega^{2,0})^{1/2}$.
Let us now review the index of the Dolbeault complex.
We will compute the index of the Dolbeault complex
on Taub-NUT space by applying the Atiyah-Bott fixed point formula.
Taub-NUT space is holomorphically isomorphic to flat $\mathbb C^2$
with local coordinates $(z_1,z_2)$,
for which the $U(1)\times U(1)$-equivariant index of the (untwisted)
Dolbeault complex
is given by
\begin{eqnarray}
\text{ind}(\bar\partial)
=
\frac{t_1 t_2}{(1-t_1)(1-t_2)}\,.
\label{index-Dolbeault}
\end{eqnarray}
Let us denote by $ U(1)_{J+R}$ the group generated
by $J_3+I_3$, the simultaneous spatial and R-symmetry rotations.
The action of $(t_1, t_2)$ on $\mathbb C^2$
is standard, $(z_1,z_2)\mapsto (t_1 z_1, t_2 z_2)$,
and is related to $U(1)_K\times U(1)_{J+R}$ as
\begin{eqnarray}
t_1=e^{-2\pi i\nu+\pi i\lambda }\,,~~~~~t_2=e^{2\pi i\nu+\pi i\lambda}\,,
\label{tnuep}
\end{eqnarray}
as can be seen from (\ref{z1z2}).
Here $e^{2\pi i\nu}$ parametrizes $U(1)_K$,
while $2\pi \lambda$ is
the angle of rotation
along the 3-axis of $\mathbb R^3$,
which is the base of the circle fibration in Taub-NUT space
(\ref{TN-metric}).
The $SU(2)$ R-symmetry action on the fields is also parametrized by $\lambda$.
For our purposes
the best way to understand the formula
(\ref{index-Dolbeault})
is to consider the group action
on the basis of sections.
For example an element of $\Omega^{0,0}$
can be expanded as
\begin{eqnarray}
\sum_{
k,l,m,n
}
c_{klmn}
z_1^{k}
\bar z_1^{l}
z_2^{m} \bar z_2^{n} \,,
\end{eqnarray}
where $k,l,m,n\in \mathbb Z_\geq 0$ and the coefficients transform as
$c_{klmn}\mapsto t_1^{-k+l} t_2^{-m+n} c_{klmn}$.
Elements of $\Omega^{0,1}$ and
$\Omega^{0,2}$
admit similar expansions.
Summing up the weights with appropriate signs determined by the degrees
in the complex, we obtain
\begin{eqnarray}
\text{ind}_\delta(\bar\partial)
&=&\sum_{k,l,m,n\geq 0}
(1-t_1-t_2+t_1 t_2)
t_1^{-k+l} t_2^{-m+n}
\nonumber\\
&=&
\frac{(1-t_1)(1-t_2)}
{
(1- e^{-\delta} t_1^{-1})(1- e^{-\delta} t_1)
(1-e^{-\delta} t_2^{-1})(1-e^{-\delta} t_2)
}\,.
\label{index-Dolbeault-reg}
\end{eqnarray}
Factors $e^{-\delta}$ with small $\delta>0$
are inserted to keep track of how we expand the numerator.
We obtain (\ref{index-Dolbeault})
from the regularized index (\ref{index-Dolbeault-reg})
by taking the limit $\delta \rightarrow 0$.
Including the gauge group action, we obtain
the index for the Dolbeault operator twisted by $R(E)$
\begin{equation}
\text{ind}_\delta(\bar D_R)
=
\frac{(1-t_1)(1-t_2)}
{
(1- e^{-\delta} t_1^{-1})(1- e^{-\delta} t_1)
(1-e^{-\delta} t_2^{-1})(1-e^{-\delta} t_2)
}
\sum_{w\in R} e^{2\pi i w\cdot a}
\,.
\end{equation}
The relationships of the self-dual and Dirac complexes
to the Dolbeault complex described above imply
that
\begin{eqnarray}
\text{ind}_\delta(D_{\text{SD},\mathbb C})
&=&(1+t_1^{-1}t_2^{-1})
\text{ind}_\delta(\bar D_\text{adj})\,,
\\
\text{ind}_\delta(D_{\text{Dirac},R})
&=&t_1^{-1/2}t_2^{-1/2}
\text{ind}_\delta(\bar D_R)\,.
\end{eqnarray}
Furthermore, the indices
of the Bogomolny and Dirac-Higgs complexes are obtained
by taking the $U(1)_K$-invariant parts.
This can be implemented by substituting
(\ref{tnuep}) and $a\rightarrow a+B\nu$ and
then integrating over $\nu$:
\begin{eqnarray}
\text{ind}(D_{\text{Bogo},\mathbb C})
&=&
\lim_{\delta \rightarrow 0}
\int_0^{1}d\nu
\left.
\text{ind}_\delta(D_{\text{SD},\mathbb C})\right|_{a\rightarrow a+B\nu}
\,,
\\
\text{ind}(D_{\text{DH},R})
&=&
\lim_{\delta \rightarrow 0}
\int_0^{1}d\nu
\left.
\text{ind}_\delta(D_{\text{Dirac},R})\right|_{a\rightarrow a+B\nu}
\,.
\end{eqnarray}
The factors $e^{-\delta}$ in the integrands
specify which poles to pick in the contour integrals.
We also need to take into account the Fourier modes on
$S^1$ that give rise to an infinite sum $\sum_n e^{in\varepsilon}$.
The formal parameter $\varepsilon$ for time translation
should be set to one at the end of the calculation.
Finally, the one-loop determinant $Z^\text{vm}_\text{1-loop}$
for the vector multiplet
is obtained by the rule $\sum_j c_j e^{w_j}
\rightarrow \prod_j w_j^{c_j}$
from
\begin{equation}
\text{ind}(D^\text{vm})
=\frac 1 2
\sum_{n\in \mathbb Z}e^{2\pi i n \varepsilon}
\text{ind}(D_{\text{Bogo},\mathbb C}) \,.
\label{ind-Dvm}
\end{equation}
The factor of $1/2$ in (\ref{ind-Dvm}) accounts for
the complexification of the Bogomolny complex.
For the hypermultiplet,
the one-loop determinant $Z_\text{1-loop}^\text{hm}$
arises if the same rule is applied to \cite{Gomis:2011pf}
\begin{equation}
\text{ind}(D^\text{hm}_R)
=
-\frac{1}{2}
\sum_{n\in \mathbb Z}e^{2\pi i n \varepsilon}
\sum_{f=1}^{N_\text{F}}
\Big(
e^{-2\pi im_f}
\text{ind}(D_{\text{DH},R})
+
e^{2\pi im_f}
\text{ind}(D_{\text{DH},R})|_{a\rightarrow-a}
\Big)
\,.
\label{ind-Dhm}
\end{equation}
Let us explain the meaning of this expression (\ref{ind-Dhm}).
The precise flavor symmetry of a massless theory is best described
in terms of half-hypermultiplets.
If an irreducible representation $R$ is real,
half-hypermultiplets can only appear in an even number $2N_\text{F}$,
and the flavor symmetry $G_\text{F}$ is $Sp(2N_\text{F})$.
The symplectic group $Sp(2N_\text{F})$ has rank $N_\text{F}$ in our convention.
For a complex irreducible representation $R$,
half-hypermultiplets always appear in conjugate pairs $R\oplus \bar R$.
With $N_\text{F}$ such pairs, the flavor symmetry is $U(N_\text{F})$.
When an irreducible representation $R$ is pseudo-real,
the theory is anomalous unless an even number $2N_\text{F}$ of half-hypermultiplets
are present \cite{Witten:1982fp}.
The flavor symmetry group in this case is $SO(2N_\text{F})$.
Parameters $m_f$ in (\ref{ind-Dhm}) are the equivariant parameters
for the flavor group $G_\text{F}$ of the massless theory,
and are related to the physical masses $M_f$
and the flavor chemical potentials $\mu_f$ as
\begin{equation}
m_f=-\mu_f+iR M_f\,.
\end{equation}
The particular combination of terms in (\ref{ind-Dhm}) was derived
in \cite{Gomis:2011pf} based on Higgsing which produces various
types of matter representations.
The indices $\text{ind}(D_{\text{Bogo},\mathbb C})$
and $\text{ind}(D_{\text{DH},R})$ were computed in \cite{Gomis:2011pf}:
\begin{eqnarray}
\text{ind}(D_{\text{Bogo},\mathbb C})
\hspace{-2mm}
&=&
\hspace{-2mm}
-
\frac{e^{\pi i\lambda}+e^{-\pi i\lambda }}2
\sum_{\alpha}
e^{2\pi i\alpha\cdot a}
\left(
e^{(|\alpha\cdot B|-1) \pi i\lambda}
+
e^{(|\alpha\cdot B|-3)\pi i\lambda}
+\ldots
+
e^{-(|\alpha\cdot B|-1)\pi i\lambda}
\right)\,,
\nonumber
\\
\text{ind}(D_{\text{DH},R})
&=&
- \frac 1 2
\sum_{w\in R}
e^{2\pi i w\cdot a}
\left(
e^{(|w\cdot B|-1)\pi i \lambda}
+
e^{(|w\cdot B|-3)\pi i \lambda}
+
\ldots
+
e^{-(|w\cdot B|-1)\pi i \lambda}
\right)
\,.
\end{eqnarray}
By applying the rule to (\ref{ind-Dvm}) and (\ref{ind-Dhm}),
we find the one-loop determinant
\begin{eqnarray}
&&\prod_{n\in \mathbb Z}\prod_{\alpha} \prod_{k=0}^{|\alpha \cdot B|-1}
\left[
n \varepsilon+\frac 1 2 \lambda
+\alpha \cdot a
+\left(\frac{|\alpha\cdot B|-1}2-k\right)\lambda
\right]^{-1/2}
\nonumber\\
&\sim&
\prod_{\alpha>0}
\prod_{k=0}^{|\alpha\cdot B|-1}
\prod_\pm
\sin^{-1/2}\left[\pi\left(
\alpha\cdot a
\pm\left(\frac{|\alpha\cdot B|}2-k\right)\lambda
\right)\right]
\nonumber\\
&=:& Z_\text{1-loop}^\text{vm}(a,\lambda;B)
\,,
\label{1-loop-vm}
\end{eqnarray}
for the vector multiplet and
\begin{eqnarray}
&&
\prod_{n\in \mathbb Z}\prod_{f=1}^{N_\text{F}}\prod_{w\in R}
\prod_{k=0}^{|w\cdot B|-1}
\left[n \varepsilon+ w\cdot a-m_f
+\left(\frac{|w\cdot B|-1}2-k\right)\lambda
\right]^{1/2}
\nonumber\\
&\sim &
\prod_{f=1}^{N_\text{F}}\prod_{w\in R}
\prod_{k=0}^{|w\cdot B|-1}
\sin^{1/2}
\left[
\pi\left(
w\cdot a-m_f +\left(\frac{|w\cdot B|-1}2 -k\right)\lambda
\right)
\right]
\nonumber\\
&=:& Z^\text{hm}_\text{1-loop}(a,m_f,\lambda;B)
\label{1-loop-hm}
\end{eqnarray}
for the hypermultiplet.
In the final expressions we set $\varepsilon$ to one.
When there is more than one matter irreducible representation
we need to take a product over them.
Combining the vector multiplet and hypermultiplet contributions,
the one-loop factor is given by
\begin{equation}
Z_\text{1-loop}(a,m_f,\lambda;B)
:=
Z_\text{1-loop}^\text{vm}(a,\lambda;B)
Z_\text{1-loop}^\text{hm}(a,m_f,\lambda;B)\,.
\label{1-loop-total}
\end{equation}
\section{Contributions from monopole screening}
\label{sec:mono}
In this section we calculate the contributions from non-perturbative
saddle points of the localization action $Q\cdot V$.
Since the bosonic part of $Q\cdot V$ is given by $||Q\cdot \Psi||^2$,
these saddle points are the solutions of the equation $Q\cdot \Psi=0$.
As we saw in Section \ref{sec:loc-eq}, the solutions of $Q\cdot \Psi=0$ are
the fixed points of the Bogomolny equations with a prescribed singularity.
\subsection{Definition of $Z_\text{mono}$}
The moduli space of the solutions of the Bogomolny equations
with a singularity prescribed by $B$ has infinitely many components.
For example, even for $B=0$ there exist the components
whose elements are smooth monopoles with charges labeled by all
$v\in\Lambda_{cr}$.
In our localization calculation only the components that contain
fixed points of the $U(1)_{J+R}\times T$-action are relevant,
where $T$ is the maximal torus of the gauge group.
Invariance under $U(1)_{J+R}\times T$-action is a strong constraint,
because the $T$-invariance for generic $a\in \mathfrak t$
requires the adjoint fields to be Abelian, {\it i.e.}, that they belong
to $\mathfrak t$.
The only Abelian solutions to the Bogomolny equations
are the singular Dirac monopole solutions,
and the singularity must be located at the point where
the 't Hooft operator is inserted.
This argument almost shows that the background configuration
(\ref{thooft-background}) is the only saddle point of the path
integral.
Abelian solutions of the Dirac form (\ref{thooft-background}),
where $B$ is replaced by some other coefficient $v\in \Lambda_{cr}+B$,
can however arise as a limit in the family of solutions
whose singularity has coefficient $B$ \cite{Kapustin:2006pk}.
Such solutions represent smooth monopoles that approach
the singular monopole and screen its charge.
See \cite{Cherkis:2007jm} for an explicit example.
For our calculation we only need to consider the components
of the moduli space that contain such solutions.
Under Kronheimer's correspondence, mentioned in Section \ref{sec:calc-index}
and reviewed in Appendix \ref{sec:mono-inst},
the Abelian solution specified by $v$ uplift to a
small instanton located at the point on Taub-NUT space where
the $S^1$ fiber degenerates.
Since our calculation needs only the local behavior of the fields
near this point, we can replace Taub-NUT space by $\mathbb C^2$.
A more satisfying justification for this replacement is
the fact that
such a small instanton solution belongs to
a component of the instanton moduli space
that is isomorphic as a complex variety
to a component of the instanton moduli space for $\mathbb C^2$
\cite{Cherkis:2008ip}.
See also \cite{Witten:2009xu}.
We denote by $\mathcal M(B,v)$ the moduli space for the Bogomolny equations
that descend from the component of the instanton moduli space.
A generic point of
$\mathcal M(B,v)$ is a solution that approaches
the background (\ref{thooft-background}) near the origin,
and the same expression with $B$ replaced by $v$ asymptotically at
infinity.
It can be shown that we need $||v||\leq ||B||$
for $\mathcal M(B,v)$ to be non-empty \cite{Gomis:2011pf}.
Since
all the fixed points in $\mathcal (B,v)$ take the form of the 't Hooft background (\ref{thooft-background})
except that $B$ is replaced by $v$,
each contributes a factor $e^{-S_\text{cl}(v)}$
computed in Section \ref{sec:classical}.
This classical contribution depends only on $v$
and is universal among the fixed points in $\mathcal M(B,v)$.
We also need to include the fluctuation determinant
$ \prod_j w_j^{c_j}$ from each fixed point, which
can be computed from the indices
of the Bogomolny and Dirac-Higgs complexes via the rule
$\sum_j c_j e^{w_j}\rightarrow \prod_j w_j^{c_j}$,
as in the one-loop analysis in Section \ref{sec:one-loop}.
By factoring out $Z_\text{1-loop}(v)$ that was computed in
Section \ref{sec:one-loop},
we denote the sum of such determinants by
\begin{equation}
Z_\text{1-loop}(v) Z_\text{mono}(B,v)
\equiv
\mathop{\sum_\text{fixed points}}_{\text{in }\mathcal M(B,v)} \prod_j w_j^{c_j}\,.
\end{equation}
This equation defines $Z_\text{mono}(B,v)$ as a function of
$B, v, a, b, m_f$, and $\lambda$.
As mentioned in footnote \ref{foot:Nek},
$\langle L\rangle$ may be thought of as a dimensional reduction
of the five-dimensional instanton partition function
with an operator insertion.
Thus $Z_\text{mono}(B,v)$ can be interpreted in terms of appropriate
characteristic classes on $\mathcal M(B,v)$.
\subsection{Monopole moduli space for $G=U(N)$}
In order to compute $Z_\text{mono}(B,v)$ explicitly,
we need a method to describe the component $\mathcal M(B,v)$
of the monopole moduli space and their fixed points.
Let us now review the ADHM construction of $\mathcal M(B,v)$ in
the case $G=U(N)$ \cite{Kapustin:2006pk}.
We consider the flat space $\mathbb C^2$
parametrized by coordinates $z=(z_1,z_2)$.
Let us set $W:=\mathbb C^N$ and $V:=\mathbb C^k$.
The instanton bundle over $\mathbb C^2$ with instanton number $k$
is described by a family of complexes
\begin{equation}
V \stackrel{\alpha(z)}\longrightarrow \mathbb C^2\otimes V\oplus W
\stackrel{\beta(z)}\longrightarrow V\,,
\end{equation}
where the maps depend on $z$ as
\begin{equation}
\alpha(z)=
\begin{pmatrix}
z_2-B_2\\
-z_1+B_1\\
-J
\end{pmatrix}
\,,
\quad
\beta(z)=
\begin{pmatrix}
z_1-B_1&z_2-B_2&-I
\end{pmatrix}
\,.
\end{equation}
When the complex ADHM equation
\begin{equation}
[B_1,B_2]+IJ=0
\end{equation}
which is equivalent to $\beta(z)\alpha(z)=0$
is satisfied,
the cohomology groups
\begin{equation}
H^0_z=\text{Ker} [\alpha(z)]\,,\quad
H^1_z=\text{Ker}[\beta(z)]/\text{Im}[\alpha(z)]\,,\quad
H^2_z=V/\text{Im} [\beta(z)]
\end{equation}
can be defined.
If $H^0_z=H^2_z=0$, $E_z=H^1_z$ describes the fiber of a smooth irreducible
instanton bundle over $\mathbb C^2$.
We are also interested in singular configurations that arise as a limit
of smooth ones, therefore we set
$E_z=H^1_z-H^0_z-H^2_z$ in general.
The Euler characteristic $\dim H^0_z-\dim H^1_z+\dim H^2_z=-\dim E_z=-N$
is independent of $z$.
A monopole solution in $\mathcal M(B,v)$ descends from
a $U(1)_K$-invariant instanton.
The group acts geometrically on $(z_1,z_2)$ as
$(z_1,z_2)\mapsto (e^{-2\pi i\nu}z_1, e^{2\pi i \nu}z_2)$ as
in
(\ref{tnuep}).
Since $(B_1,B_2)$ represent the positions of the instantons,
they transform as
$(B_1,B_2)\mapsto (e^{-2\pi i\nu}B_1, e^{2\pi i \nu}B_2)$.
The group $U(1)_K$ also acts on the gauge bundle.
The fiber $E_0$ at $z=0$ if mapped to itself,
and its character for $U(1)_K$ is given by $e^{2\pi i B\nu}$
where $e^{2\pi \nu}\in U(1)_K$ and the charge $B$
of the 't Hooft operator is regarded
as a $N\times N$ diagonal matrix.
The group $U(1)_K$ also acts on $W$ and $V$.
Since $W$ represents the fiber $E_\infty$ at $z=\infty$,
its character is ${\rm Tr} e^{2\pi i v \nu}$.
The character of $V$ can be written as $e^{2\pi i K\nu}$
with a $k\times k$ diagonal matrix $K$.
The identification of $E_{z}$
with $H^1_z-H^0_z-H^2_z$ implies that $K$ is determined by\footnote{A warning on notation. The ``$K$'' in $U(1)_K$ stands for Kronheimer.
The matrix $K$ is the weight of $U(1)_K$ acting on the $k$-dimensional
vector space on which $B_1$ and $B_2$ act as endomorphisms.
}
\begin{equation}
{\rm Tr} e^{2\pi i B\nu}
={\rm Tr} e^{2\pi i v\nu}+(e^{2\pi i\nu}+e^{-2\pi i\nu}-2) {\rm Tr} e^{2\pi i K\nu}
\quad
\quad e^{2\pi i\nu} \in U(1)_K
\label{B-v-K-condition-2}
\end{equation}
up to conjugation.
To describe $\mathcal M(B,v)$, we impose $U(1)_K$ invariance on the ADHM data.
Namely the ADHM data must satisfy the conditions
\begin{equation}
- B_1+[K,B_1]=0\,,\quad
B_2+[K,B_2]=0\,,\quad
KI-Iv=0\,,\quad
v J-J K=0\,.
\label{U(1)K-inv-eq}
\end{equation}
For the instanton moduli space, one would take a quotient by $GL(k,\mathbb C)$.
The matrix $K$ breaks the $GL(k,\mathbb C)$ into its commutant
$\prod_rGL(k_r,\mathbb C)$.
Two combinations of such data are considered equivalent if they are related by an action of $\prod_r GL(k_r,\mathbb C)$:
\begin{equation}
(B_1, B_2, I,J)
\sim
(g B_1 g^{-1}, g B_2 g^{-1}, g I,J g^{-1})
\quad \quad
g\in \prod_r GL(k_r,\mathbb C)\,.
\end{equation}
Thus the complex variety $\mathcal M(B,v)$ is given by
the holomorphic quotient
\begin{eqnarray}
\mathcal M(B, v)=
\left \{
(B_1,B_2,I,J)
\left|
\begin{array}{ccc}
- B_1+[K,B_1]&=&0\\
B_2+[K,B_2]&=&0\\
K I - I M&=&0\\
M J - J K&=&0\\
\end{array}
\right.
\right\}
{\Big /}
\prod_r GL(k_r,\mathbb C)\,.
\end{eqnarray}
The notion of fixed points requires a regularization
of singularities in $\mathcal M(B,v)$.
In this paper we do not attempt to describe the regularization
in detail though we believe that this is important for
the precise definition of the 't Hooft loop with a given magnetic charge
$B$.
See Section \ref{sec:discussion} for a further discussion
on this point.
We will use a partial regularization that descends from the moduli space of non-commutative instantons that smooth the small instanton singularities.
This led to a prescription, based on contour integrals,
for how to take into account the fixed point contributions in \cite{Gomis:2011pf}.
Here we give an alternative prescription
for the calculation
of the fixed points and their contributions.
\subsection{Fixed points and their contributions}
Next we turn to the description of fixed points.
We need to know which fixed point $\vec Y$ on the instanton
moduli space descends to
the specific component $\mathcal M(B,v)$ of the monopole moduli space.
The fixed points are given by the ADHM data $(B_1,B_2,I,J)$
that satisfy
\begin{equation}
\begin{aligned}
\varepsilon_1 B_1+[\phi, B_1]=0\,,
\quad
\varepsilon_2 B_2+[\phi, B_2]=0\,,
\\
\phi I-I a=0\,,
\quad
(\varepsilon_1+\varepsilon_2) J+ a J- J\phi=0\,.
\end{aligned}
\label{inst-fixed-point-eq}
\end{equation}
for any $(\varepsilon_1,\varepsilon_2,a)\in \text{Lie}\left[U(1)\times U(1)\times T\right]$
for some $\phi=\text{diag}(\phi_1,\ldots,\phi_k)$
parametrizing the Cartan subalgebra
of $\prod_r U(k_r)\subset U(k)$.
Solutions to these equations are known
\cite{Nekrasov:2002qd, Nakajima:2003pg}
and are expressed in terms of Young diagrams $\vec Y$.
See \cite{Bruzzo:2002xf} for explicit expressions for $(B_1, B_2, I,J)$
at the fixed point $\vec Y$.
Here we only need the expressions\footnote{In this subsection we use Greek alphabets $\alpha,\beta,\ldots$
to denote the $U(N)$ indices,
and use $(i,j)$ to denote the location of a box
in a Young diagram.
}
for $\phi_s$ \cite{Nekrasov:2002qd}
\begin{equation}
\phi_s= (i_s-1)\varepsilon_1+(j_s-1)\varepsilon_2+a_{\alpha(s)}
\text{ where $\alpha(s)\in \{1,\ldots, N\}$ is such that } s\in Y_{\alpha(s)}\,.
\label{phi-fixed-point}
\end{equation}
Since the fixed point $\vec Y$ in the instanton moduli space
satisfy the general $U(1)^2\times T$-invariance condition
(\ref{inst-fixed-point-eq}) together with (\ref{phi-fixed-point}),
it also satisfies the $U(1)_K$ invariance condition
(\ref{U(1)K-inv-eq}) if $U(1)_K$ is embedded in
$U(1)^2\times U(N)$ in such a way that their actions are compatible.
Since the embedding is given by the substitution
\begin{equation}
\varepsilon_1\rightarrow -\nu,,
\quad
\varepsilon_2\rightarrow \nu\,,
\quad
a \rightarrow a+v\nu\,,
\quad
\end{equation}
the $U(1)_K$-invariant fixed points correspond to
\begin{equation}
\vec Y \quad \text{ such that} \quad
K_s= v_{\alpha(s)}+j_{\alpha(s)}-i_{\alpha(s)}
\label{vec-Y-cond}
\end{equation}
up to a permutation of $s\in \{1,\ldots, k\}$.
To obtain the weights $w_j$ each fixed point
contributes, we can combine the method in
Section \ref{sec:one-loop} with the known result
for the Dolbeault index at the fixed point.
We recall from that section that
the Dolbeault index on $\mathbb C^2$,
defined by a formal application of the Atiyah-Bott formula,
is given by
\begin{equation}
\text{ind}(\bar D_\text{adj})
=
\sum_{\alpha,\beta=1}^N e_\alpha e_\beta^{-1} \frac{1}{(1-t_1^{-1})
(1-t_2^{-1})} \,,
\end{equation}
where $e_\alpha = e^{2\pi i a_\alpha}$.
Let us define
\begin{equation}
\chi(Y) = \sum_{(i,j) \in Y} t_1^{i-1} t_2 ^{j-1}
\end{equation}
and the conjugate
\begin{equation}
\chi(Y)^* = \sum_{(i,j) \in Y} t_1^{1 - i} t_2 ^{1 - j}\,.
\end{equation}
Then the local index for the Dolbeault operator at the fixed point $\vec Y$ is
given by
\begin{equation}
\begin{aligned}
\text{ind}(\bar D_\text{adj})_{\vec Y}
=& \sum_{\alpha,\beta = 1}^{N} e_\alpha e_\beta^{-1}
\left ( \frac{ 1}{(1 - t_1)(1- t_2)} - \chi(Y_{\alpha}) \right)
\\
&
\quad\quad
\times
\left ( \frac{ 1}{(1 - t_1^{-1})(1- t_2^{-1})} - \chi(Y_\beta)^*
\right)(1-t_1)(1-t_2)\,.
\end{aligned}
\label{eq:character}
\end{equation}
As shown in Section \ref{sec:calc-index},
the one-loop determinant is obtained from
the non-polynomial part $\text{ind}(\bar D_\text{adj})^\text{1-loop}\equiv
\sum_{\alpha,\beta} e_\alpha e_\beta^{-1} (1-t_1^{-1})^{-1}(1-t_2^{-1})^{-1}
$
of (\ref{eq:character}).
The rest of (\ref{eq:character}) is a Laurent polynomial, which we denote by
$\text{ind}(\bar D_\text{adj})_{\vec Y}^\text{inst}$.
It is nothing but (minus) the character of the tangent space to
the moduli space, and can be rewritten as \cite{Nakajima:2003pg}
\begin{equation}
\text{ind}(\bar D_\text{adj})_{\vec Y}^\text{inst}
=-\sum_{\alpha,\beta = 1}^{N} e_\alpha e_\beta^{-1}
\left(
\sum_{s\in Y_\alpha} t_1^\text{$-L_{Y_\beta}(s)$} t_2^{A_{Y_\alpha}(s)+1}
+\sum_{t\in Y_\beta}
t_1^{L_{Y_\alpha}(t)+1} t_2^\text{\raisebox{.5mm}[0mm][0mm]{$-A_{Y_\beta}(t)$}}
\right)\,.
\end{equation}
We have introduced the arm- and leg-lengths
\begin{equation}
A_Y(s)=\lambda_i-j\,,\quad L_Y(s)=\lambda^T_j-i\,,
\end{equation}
where $\lambda_i$ and $\lambda^T_i$ are the numbers of boxes in the $i$-th row
and column of $Y$, respectively.
Let us denote
the $U(1)_K$-invariant part of
$\sum_{n\in\mathbb Z} e^{2\pi in\varepsilon}
((1+t_1^{-1}t_2^{-1})/2) \text{ind}(\bar D_\text{adj})^{\text{1-loop}}$
by $\text{ind}(D^\text{vm})^\text{1-loop}$,
where $U(1)_K$ acts on the gauge bundle with a generator $v$.
It gives rise to
$Z_\text{1-loop}^\text{vm}(v)$
via the rule $\sum_j c_j e^{w_j}\rightarrow
\prod_j w_j^{c_j}$ as in Section \ref{sec:one-loop}.
Similarly we define
\begin{equation}
\text{ind}(D^\text{vm})^\text{mono}_{\vec Y}
\equiv \text{ $U(1)_K$-invariant part of }
\sum_{n\in\mathbb Z} e^{2\pi in\varepsilon}
\frac{1+t_1^{-1}t_2^{-1}}{2} \text{ind}(\bar D_\text{adj})^\text{inst}_{\vec Y}\,.
\end{equation}
The same rule applied to this gives a contribution to
$Z_\text{mono}^\text{vm}(B,v)$.
The $U(1)_K$-invariant terms arise from the triples
\begin{equation}
(\alpha, \beta, s\in Y_\alpha)
\quad \text{ such that }
\quad
v_\alpha-v_\beta+L_{Y_\beta}(s)+A_{Y_\alpha}(s)+1=0
\label{alpha-beta-s-cond}
\end{equation}
which contribute
\begin{equation}
-\frac 1{2}
\sum_{n\in \mathbb Z} e^{2\pi in\varepsilon}
(1+e^{-2\pi i\lambda})
e^{2\pi i(a_\alpha-a_\beta)} e^{\pi i(A_{Y_\alpha}(s)-L_{Y_\beta}(s)+1) \lambda}\,,
\end{equation}
and also from the triples
\begin{equation}
(\alpha, \beta, t\in Y_\beta) \quad \text{ such that }
\quad
v_\alpha-v_\beta-L_{Y_\alpha}(t)-A_{Y_\beta}(t)-1=0
\label{alpha-beta-t-cond}
\end{equation}
which contribute
\begin{equation}
-\frac 1{2}
\sum_{n\in \mathbb Z} e^{2\pi in\varepsilon}
(1+e^{-2\pi i\lambda})
e^{2\pi i(a_\alpha-a_\beta)} e^{\pi i(L_{Y_\alpha}(s)-A_{Y_\beta}(s)+1) \lambda}
\end{equation}
to $\text{ind}(D^\text{vm})_{\vec Y}^\text{mono}$.
By applying the rule $\sum_j c_j e^{w_j}\rightarrow
\prod_j w_j^{c_j}$,
we find the vector multiplet contribution to $Z_\text{mono}^\text{vm}$
\begin{equation}
\begin{aligned}
z^\text{vec}_{\vec Y}=
&
\prod_{(\alpha, \beta, s)}
\prod_\pm
\left(\sin\left[
\pi
\left(
a_\alpha-a_\beta+\frac{1}{2}(A_{Y_\alpha}(s)-L_{Y_\beta}(s)\pm 1
)\lambda \right)
\right]\right)^{-1}
\,.
\label{mono-vec}
\end{aligned}
\end{equation}
We emphasize that
the products are over the triples $(\alpha, \beta, s)$
satisfying
(\ref{alpha-beta-s-cond}).
The contributions from $(\alpha,\beta,t)$ in
(\ref{alpha-beta-t-cond}) are identical to those from $(\alpha,\beta,s)$
in (\ref{alpha-beta-s-cond}).
Thus the power in (\ref{mono-vec}) is $-1$, not $-1/2$.
The same remark applies to (\ref{mono-adj}) below.
For a single hypermultiplet in the adjoint representation,
we need to consider
\begin{equation}
\begin{aligned}
\text{ind}(D^\text{hm}_\text{adj})_{\vec Y}^\text{mono}
&\equiv \text{$U(1)_K$-invariant part of }
\\
&\quad\quad
-\frac{e^{2\pi im}+ e^{-2\pi im}}{2}
\sum_{n\in\mathbb Z} e^{2\pi in\varepsilon}\,
t_1^{-1/2}t_2^{-1/2} \text{ind}(\bar D_\text{adj})_{\text{inst}}\,.
\end{aligned}
\end{equation}
From this we get the contribution of an adjoint hypermultiplet
\begin{equation}
\begin{aligned}
z^\text{adj}_{\vec Y}=
&
\prod_{(\alpha, \beta, s)}
\prod_\pm
\sin\left[
\pi
\left(
a_\alpha-a_\beta+\frac{1}{2}(A_{Y_\alpha}(s)-L_{Y_\beta}(s)
)\lambda
\pm m \right)
\right]
\,.
\end{aligned}
\label{mono-adj}
\end{equation}
The product is over the same triples as above.
For a hypermultiplet in the fundamental representation,
we need the Dolbeault index for the corresponding bundle\footnote{When used in instanton counting, this
leads to the contribution of a fundamental hypermultiplet.
}
\begin{equation}
\begin{aligned}
\text{ind}(\bar D_\text{fund})_{\vec Y}
=& \sum_{\alpha = 1}^{N} e_\alpha t_1 t_2
\left ( \frac{ 1}{(1 - t_1)(1- t_2)} - \chi(Y_{\alpha}) \right) \,.
\end{aligned}
\label{eq:character-fund}
\end{equation}
Thus the Dirac index is
\begin{equation}
\begin{aligned}
\text{ind}(\bar D_\text{Dirac,fund})_{\vec Y}
=& \sum_{\alpha = 1}^{N} e_\alpha t_1^{1/2} t_2^{1/2}
\left ( \frac{ 1}{(1 - t_1)(1- t_2)} - \chi(Y_{\alpha}) \right) \,.
\end{aligned}
\label{eq:character-fund2}
\end{equation}
For a pair
\begin{equation}
(\alpha, s\in Y_\alpha)
\quad \text{ such that}
\quad
v_\alpha-i_s+j_s=0\,,
\label{alpha-s-cond}
\end{equation}
it contributes
\begin{equation}
- e^{2\pi i a_\alpha} e^{\pi i(i_s+j_s-1)\lambda}
\end{equation}
to $\text{ind}(D_\text{DH,fund})_{\vec Y}^\text{mono}$ and thus
\begin{equation}
\frac 1 2 \sum_{n\in \mathbb Z} e^{2\pi in \varepsilon}
(
e^{2\pi i a_\alpha} e^{\pi i(i_s+j_s-1)\lambda}e^{-2\pi i m}
+
e^{-2\pi i a_\alpha} e^{-\pi i(i_s+j_s-1)\lambda}e^{2\pi i m}
)
\end{equation}
to $\text{ind}(D^\text{hm}_\text{fund})_{\vec Y}^\text{mono}$.
Then
\begin{equation}
z_{\vec Y}^\text{fund}
(a,m,\lambda;B,v)
=
\prod_{(\alpha,s)}
\sin\left[\pi\left(
a_\alpha -m
+ \frac{1}{2}\left( i_s+j_s-1
\right)\lambda
\right)\right]\,.
\label{mono-fund}
\end{equation}
Again we stress that the product is over the pairs
$(\alpha,s)$ satisfying
(\ref{alpha-s-cond}).
The total monopole screening contribution is then given by
\begin{equation}
Z_\text{mono}(a,m_f,\lambda;B,v)
=\sum_{\vec Y} z_{\vec Y}^\text{vec}(a,\lambda;B,v)
\prod_R\prod_f z_{\vec Y}^R(a,m_f;B,v)\,,
\label{mono-total}
\end{equation}
where the sum is over $N$-tuples of Young diagrams $\vec Y$
satisfying (\ref{vec-Y-cond}), and the product is over the
matter representations $R$.
Explicit expressions for $Z_\text{mono}(B,v)$ will appear
as part of the operator vevs
in Section \ref{sec:gauge-results}.
\section{Gauge theory results}
\label{sec:gauge-results}
For a Wilson operator in an arbitrary representation $R$,
the on-shell action vanishes.
The only saddle point in the path integral is the trivial one,
and the one-loop determinant is $1$ due to Bose-Fermi cancellations.
Thus the expectation value is given by evaluating the holonomy
(\ref{wilson-def}) in the background:
\begin{equation}
\langle W_R\rangle={\rm Tr}_R \exp\left[
2\pi i R \left(A_\tau^{(\infty)}+i\Phi_0^{(\infty)}\right)
\right]
={\rm Tr}_R e^{2\pi i a}\,,
\end{equation}
where $a$ was defined in (\ref{def-ab}).
For the 't Hooft operator, we combine the classical, one-loop,
and monopole screening contributions from the previous sections:
\begin{equation}
\langle T_B\rangle= \sum_{v} e^{2\pi i v\cdot b}
Z_\text{1-loop}(v)
Z_\text{mono}(B,v)\,.
\end{equation}
\subsection{$SU(2)$ $\mathcal N=2^*$}
For $SU(2)$, it is convenient to substitute
\begin{equation}
a\rightarrow\left(
\begin{array}{cc}
a&\\
&-a
\end{array}
\right)
\,,
\quad\quad
b\rightarrow\left(
\begin{array}{cc}
b&\\
&-b
\end{array}
\right)
\end{equation}
with the understanding that in the following the symbols $a$ and $b$
are complex numbers rather than matrices.
For this gauge group we can label the line operators
by a pair of integers $(p,q)$, where $p$ and $q$ are
magnetic and electric charges respectively
\cite{Kapustin:2006hi,Kapustin:2007wm,Drukker:2009tz},
and they are related to the coweight and
the highest weight of the representation as
\begin{equation}
\begin{aligned}
& B=(p/2,-p/2)\equiv \text{diag}(p/2,-p/2)\in \Lambda_{cw}\,,
\\
&(q/2,-q/2)\equiv \text{diag}(q/2,-q/2)\in \Lambda_w\quad
\leftrightarrow \quad\text{spin $q/2$ representation}\,.
\end{aligned}
\label{charge-dict}
\end{equation}
The most basic Wilson operator $W_{1/2}=L_{0,1}$
corresponding to spin $1/2$ has an expectation value
\begin{equation}
\langle W_{1/2}\rangle=
\langle L_{0,1}\rangle=
e^{2\pi i a}+e^{-2\pi i a}\,.
\label{SU22starL01}
\end{equation}
For the minimal 't Hooft operator $T_{1/2}=L_{1,0}$ that is S-dual to $W_{1/2}$,
we find
\begin{equation}
\langle T_{1/2}\rangle=
\langle L_{1,0}\rangle=
( e^{2\pi i b}+e^{-2\pi i b})
\left(
\frac{
\sin\left(2\pi a+\pi m\right)
\sin\left(2\pi a-\pi m\right)
}{
\sin\left(2\pi a+\frac{\pi}2 \lambda\right)
\sin\left(2\pi a-\frac{\pi}2 \lambda\right)
}
\right)^{1/2}\,.
\label{SU22starL10}
\end{equation}
For the minimal dyonic loops $L_{1,\pm 1}$,
\begin{equation}
\langle L_{1,\pm 1}\rangle=
( e^{2\pi i (b\pm a)}+e^{-2\pi i (b\pm a)})
\left(
\frac{
\sin\left(2\pi a+\pi m\right)
\sin\left(2\pi a-\pi m\right)
}{
\sin\left(2\pi a+\frac{\pi}2 \lambda\right)
\sin\left(2\pi a-\frac{\pi}2 \lambda\right)
}
\right)^{1/2}\,.
\label{SU22starL1pm1}
\end{equation}
The simplest example with monopole screening contribution is given by
\begin{equation}
\begin{aligned}
\langle L_{2,0}\rangle
=&
(e^{4\pi i b}
+
e^{-4\pi i b})
\frac{
\prod_{s_1,s_2=\pm 1}\sin^{1/2}\left(2\pi a+s_1 \pi m
+s_2 \frac{\pi}{2}\lambda\right)
}{
\sin^{1/2}\left(
2\pi a+\pi\lambda
\right)
\sin^{1/2}\left(
2\pi a-\pi\lambda
\right)
\sin\left(
2\pi a
\right)
}
\\
&
\quad
+
\sum_{s=\pm}
\frac{\prod_\pm\sin \pi (2a\pm m+ s\lambda/2) }
{\sin( 2\pi a )\sin \pi(2a+s \lambda)}
\,,
\end{aligned}
\label{SU22starL20}
\end{equation}
where we used (\ref{1-loop-total}) and (\ref{mono-total}).
We observe that this is the Moyal product of the minimal
't Hooft operator vev with it self,
\begin{equation}
\langle L_{2,0}\rangle
=
\langle L_{1,0}\rangle
*
\langle L_{1,0}\rangle \,.
\end{equation}
In the $SU(2)$ case $*$ is defined by
\begin{equation}
(f*g)(a,b)\equiv
e^{i\frac{\lambda}{8\pi}
(
\partial_{b} \partial_{a'}
-
\partial_a \partial_{b'}
)
}f(a,b) g(a',b')
|_{a'=a,b'=b}
\end{equation}
with a different coefficient due to the factor of 2
in the inner product
\begin{equation}
a\cdot b\rightarrow {\rm Tr}[\text{diag}(a,-a)\cdot \text{diag}(b,-b) ]=2ab\,.
\label{ab-prod}
\end{equation}
In Section \ref{sec:non-com}, we will explain how the Moyal product appears
from the structure of the path integral.
The precise choice of signs and relative numerical normalizations among terms
is difficult to fix purely in gauge theory without additional assumptions.
In the examples considered in this paper we choose to be pragmatic and
make the choice by assuming physically reasonable structures such as
Moyal multiplication, correspondence with the Verlinde operators,
as well as agreement with classical $SL(2,\mathbb C)$ holonomies in the
$\lambda \rightarrow 0$ limit.
\subsection{$U(N)$ $\mathcal N=2^*$}
For the gauge group $U(N)$, the minimal 't Hooft operators,
with charges\footnote{The Cartan subalgebra of $U(N)$ is spanned by
real diagonal matrices.
For $SU(N)$ they must be traceless.
We often drop ``diag''
in $a=\text{diag}(a_1,\ldots,a_N)$ to simplify notation.
The inner product is defined by the trace $a \cdot a'={\rm Tr}\, a a'$,
and this is used to identify the Cartan algebra with its dual.
}
$B=(\pm 1, 0^{N-1})$ (the power indicates the number of repeated entries) corresponding to the
fundamental and anti-fundamental representations of the Langlands dual group,
have the expectation values
\begin{equation}
\langle T_{B=(\pm 1, 0^{N-1})}\rangle
=
\sum_{l=1}^N
e^{\pm 2\pi i b_l}
\left(
\prod_\pm \prod_{j\neq l}
\frac{
\sin\pi(a_l-a_j\pm m)
}{
\sin\pi(a_l-a_j\pm \lambda/2)
}
\right)^{1/2}\,.
\label{UN2starBpm10N-1}
\end{equation}
For the magnetic charge $B=(1,-1,0^{N-2})$, corresponding
to the adjoint representation,\begin{equation}
\begin{aligned}
& \langle T_{B=(1,-1,0^{N-2})}\rangle
\\
=&
\sum_{k\neq l}
e^{2\pi i (b_k-b_l)}
\left[
\frac{
\left[
{\displaystyle \prod_{\pm,\pm}} \sin\pi\left( a_{kl}\pm m\pm \lambda/ 2\right)
\right]
\left[
{\displaystyle
\prod_{\pm}
\prod_{j\neq k,l}
}
\sin\pi\left( a_{kj}\pm m\right)
\sin\pi\left( a_{lj}\pm m\right)
\right]
}{
\sin^2\pi a_{kl}
{\displaystyle \prod_{\pm}}
\sin\pi\left(a_{kl}\pm \lambda\right)
\left[
{\displaystyle
\prod_{\pm}
\prod_{j\neq k,l}
}
\sin\pi\left( a_{kj}\pm \lambda/ 2\right)
\sin\pi\left( a_{lj}\pm\lambda/ 2\right)
\right]
}
\right]^{1/2}
\\
&
\quad\quad
+
\sum_{l=1}^N
\prod_{j\neq l}
\frac{
\prod_\pm
\sin\pi( a_{lj}\pm m+\lambda/2)
}{
\sin \pi a_{lj}\sin\pi (a_{lj}+\lambda)
}
\,.
\end{aligned}
\label{UN2starB1-10N-2}
\end{equation}
From (\ref{UN2starBpm10N-1}) and (\ref{UN2starB1-10N-2}) we find that
\begin{equation}
\langle T_{B=(1,-1,0^{N-2})}\rangle=
\langle T_{B=(-1,0^{N-1})}\rangle*
\langle T_{B=(1,0^{N-1})}\rangle\,.
\end{equation}
For $B=(2,0^{N-1})$,
\begin{equation}
\begin{aligned}
\langle T_{B=(2,0^{N-1})}\rangle
&=
\sum_{k=1}^N e^{4\pi i b_k}
\left(
\prod_{j\neq k}
\frac{
\prod_{\pm,\pm}
\sin\pi (a_{kj}\pm m \pm \lambda/2)
}
{
\left[
\sin^2\pi a_{kj}
\prod_\pm\sin\pi (a_{kj}\pm \lambda)
\right]
}
\right)^{1/2}
\\
&
\quad \quad \quad
+\sum_{k\neq l}
\frac{
\prod_\pm \sin\pi (a_{kl}\pm m+\lambda/2)
}
{
\sin\pi (a_{kl}+\lambda)
\sin\pi a_{kl}
}\,.
\end{aligned}
\end{equation}
For this we find
\begin{equation}
\langle T_{B=(2,0^{N-1})}\rangle
=
\langle T_{B=(1,0^{N-1})}\rangle
*
\langle T_{B=(1,0^{N-1})}\rangle\,.
\end{equation}
Results for the gauge group $SU(N)$ can be obtained by taking
$a$ and $b$ traceless.
\subsection{$U(2)$ $N_\text{F}=4$ }
For the minimal 't Hooft operator in this theory, we have
\begin{equation}
\begin{aligned}
\langle T\rangle
=&
e^{\pi i b_{12}}
\left(
\frac{
\prod_{f=1}^4 \sin\pi(a_1-m_f)\sin\pi(a_2-m_f)}
{
\sin^2\pi a_{12} \prod_\pm \sin\pi(a_{12}\pm\lambda)
}
\right)^{1/2}
\\
&+
e^{-\pi i b_{12}}
\left(
\frac{
\prod_{f=1}^4 \sin\pi(a_1+m_f)\sin\pi(a_2+m_f)}
{
\sin^2\pi a_{12} \prod_\pm \sin\pi(a_{12}\pm\lambda)
}
\right)^{1/2}
\\
&+
\frac{
\prod_{f=1}^{4}
\sin \pi
\left(
a_1-m_f +\frac{\lambda}{2}
\right)
}
{
\sin \pi a_{12}
\sin \pi \left(-a_{12}-\lambda\right)
}
+
\frac{
\prod_{f=1}^{4}
\sin \pi
\left(
a_2-m_f +\frac{\lambda}{2}
\right)
}
{
\sin \pi a_{21}
\sin \pi \left(-a_{21}-\lambda\right)
}\,.
\end{aligned}
\label{U2NF4}
\end{equation}
We have defined $a_{jk}=a_j-a_k$.
\subsection{$U(N)$ $N_\text{F}=2N$ }
\label{sec:UNNF=2N}
For the minimal 't Hooft operator given by
the magnetic charge $B=\text{diag}(1,-1,0^{N-2})$ corresponding to
the adjoint representation,
we obtain
\begin{equation}
\begin{aligned}
& \langle T_B\rangle
\\
=&
\mathop{\sum_{1\leq k,l\leq N}}_{k\neq l}
e^{\pi i(b_k-b_l)}
\frac{
\left[
\prod_{f=1}^N
\sin\pi (a_k-m_f)
\sin\pi (a_l-m_f)
\right]^{1/2}
}{
\sin\pi a_{kl}
{\displaystyle \prod_\pm }
\left[
\sin \pi (a_{kl}\pm\lambda)
{\displaystyle
\prod_{j\neq k,l}}
\sin\pi(a_{kj}\pm \lambda/2)
\sin\pi(a_{jl}\pm \lambda/2)
\right]
^{1/2}
}
\\
&\quad\quad
+
\sum_{l=1}^N
\frac{
\prod_{f=1}^{2N}
\sin \pi
\left(
a_l-m_f +\frac{\lambda}{2}
\right)
}
{
\prod_{j\neq l}
\sin \pi a_{lj}
\sin \pi\left(- a_{lj}-\lambda\right)
}\,.
\end{aligned}
\label{UNSQCD}
\end{equation}
We have introduced the notation $a_{jk}\equiv a_j-a_k$.
We emphasize that (\ref{UNSQCD}) and (\ref{U2NF4}) are the vev of the 't Hooft operator
in the $U(N)$ and $U(2)$ theories, not in the $SU(N)$ and $SU(2)$ theories.
We will compare (\ref{UNSQCD}) and (\ref{U2NF4})
with the Verlinde operators in Toda and Liouville theories
in Section \ref{sec:S4-CFT} that we will propose to
be related to the line operators in the $SU(N)$ and $SU(2)$ theories.
While we do not have a computational method intrinsic to $SU(N)$,
we will see that (\ref{UNSQCD}) and (\ref{U2NF4}), when $a$ is restricted
to be traceless, do reproduce $a$-dependent
terms in the CFT results.
\section{Noncommutative algebra and quantization}
\label{sec:non-com}
By using the structure of the path integral we have
found, in this section we show that the vevs of the line operators
on $S^1\times \mathbb R^3$, inserted on the 3-axis ($x^1=x^2=0$),
form a non-commutative algebra,
when the axis is considered as time
and the operators are time-ordered.
We will begin with the $U(1)$ case and then discuss the general
gauge group.
\subsection{Maxwell theory}
\label{sec:non-com-U1}
Let us explain how non-commutativity arises
in the algebra of Wilson-'t Hooft
operators in Maxwell theory on $S^1\times \mathbb R^3$
upon twisting by a spatial rotation along the $S^1$.
We begin with an intuitive explanation based on classical fields
\cite{Gaiotto:2010be}.
By taking $S^1$ as time,
the expectation value of the product of Wilson ($W$)
and 't Hooft ($T$) operators
can be thought of as the trace
\begin{equation}
\langle W\cdot T\rangle= {\rm Tr}_{\mathcal H(W\cdot T)} (-1)^F e^{-2\pi R H}
e^{2\pi i\lambda J_3}
\end{equation}
taken in the Hilbert space $\mathcal H(W\cdot T)$
defined by the line operators.
The space $\mathcal H(W\cdot T)$ differs from
the simple product $\mathcal H(W)\otimes \mathcal H(T)$
because when both $W$ and $T$ are present,
their electric and magnetic fields
produce the Poynting vector $\vec E\times \vec B$ that carries
a non-zero angular momentum.
The orientation of the Poyinting vector,
and therefore the phase $e^{2\pi i\lambda J_3}$,
depends on the relative
positions of the operators on the 3-axis.
Next we present an approach
suitable for localization.
For simplicity let us turn off the theta angle.
The line operator $L_{p,q}$ with magnetic and electric charges $(p,q)$
at the origin $\vec x=0$
is
defined by the path integral over the fluctuations around the
singular background
\begin{equation}
A
=
A^{(\infty)}_\tau d\tau
+p \frac{\cos\theta}2
d\varphi
\label{U1-background}
\end{equation}
with the insertion of the holonomy
\begin{equation}
e^{-i q\oint_{S^1}A}\,.
\label{U1-Wilson}
\end{equation}
We note here that the expression for the monopole field
in (\ref{U1-background})
has Dirac strings in two directions ($\theta=0,\pi$).
The expectation value $\langle L_{p,q}\rangle$
is a function of $(a,b)$, which are normalized
electric and magnetic background Wilson lines
\begin{equation}
a\equiv R A_\tau^{(\infty)}\,,
\quad\quad
b\equiv\frac{\Theta}{2\pi}\,.
\end{equation}
We claim that the path integral yields the expectation value
\begin{equation}
\langle L_{p,q}\rangle=e^{-2\pi i (q a+ p b)}\,.
\end{equation}
The magnetic part is essentially the definition of the magnetic
Wilson line $\Theta$, which is defined as the chemical potential for the
magnetic charge at infinity.
The electric part arises because the holonomy
(\ref{U1-Wilson}) is evaluated against the background Wilson line.
Let us introduce a twist along the $S^1$.
If we think of the circle as the time direction,
we can write
\begin{equation}
\langle L_{p,q}\rangle= {\rm Tr}_{\mathcal H(L_{p,q})} (-1)^F e^{-2\pi R H}
e^{2\pi i\lambda J_3}\,,
\end{equation}
where $J_3$ is the Cartan
generator of the spatial rotation group $SU(2)$.
The twist by $J_3$ means that we rotate the system by angle
$2\pi \lambda$ as we go along $S^1$,
i.e., we introduce the identification
\begin{equation}
(\tau+2\pi R, \varphi)\sim (\tau, \varphi+2\pi\lambda)\,.
\end{equation}
In terms of the new coordinates
$(\tau',\varphi')=(\tau,\varphi+
\frac{\lambda}{R}\tau
)$, the identification is simply
\begin{equation}
(\tau'+2\pi R,\varphi')\sim (\tau', \varphi'
)\,.
\end{equation}
The components of the gauge field are related as
\begin{equation}
A_{\tau'}=A_{\tau}-\frac{\lambda}{R} A_{\varphi}\,,
\quad
A_{\varphi'}=A_\varphi\,.
\end{equation}
Note that $A_\varphi$ represents a holonomy around the Dirac strings.
In our choice of local trivialization $A_\varphi(\theta=\pi/2)=0$,
so we have a simple relation
\begin{equation}
\oint_{S^1} A= 2\pi a\quad\text{ at $\quad\theta=\pi/2$.}
\end{equation}
Thus the monopole field does not contribute to the holonomy as claimed above,
in fact even after twisting.
The holonomies at $\theta\neq \pi/2$ are, however, shifted from $a$.
Indeed we find
\begin{equation}
\oint_{S^1} A= \int_0^{2\pi R} d\tau' A_{\tau'}=2\pi
\left( a \mp \frac{p}{2}\lambda\right)
\quad
\text{ at }
\quad
\theta=
\left\{
\begin{array}{lll}
0\,,\\
\pi\,.
\end{array}
\right.
\end{equation}
One can picture the shift as arising from the holonomy winding around
the Dirac strings.
Then for the product of Wilson and 't Hooft
operators $W\equiv L_{0,1}, T\equiv L_{1,0}$
\begin{equation}
W(\vec x=(0,0, z))
\cdot
T(\vec x=0)\,,
\end{equation}
its expectation value is given by
\begin{equation}
\langle W(\vec x=(0,0, z))
\cdot
T(\vec x=0)\rangle
=
e^{-2\pi i (a\mp \frac 1 2 \lambda)} e^{-2\pi i b}
\text{ for }
\left\{
\begin{array}{lll}
z>0\,,\\
z<0\,.
\end{array}
\right.
\end{equation}
The Wilson line operator (\ref{U1-Wilson}) for $z>0$
is evaluated at $\theta=0$,
and for $z<0$ at $\theta=\pi$.
The difference $\lambda$ between the shifts in $a$ at $z>0$ and $z<0$
is independent of the choice of local trivialization.
We can also see that the expectation value
of the product of operators is given by the Moyal product
of the expectation values:
\begin{equation}
\langle W(z)
\cdot
T(0)\rangle
=
\left\{
\begin{array}{lll}
\langle W\rangle *
\langle T\rangle
\quad
\text{ for $z>0$}\,,
&
\\
\langle T\rangle
*
\langle W\rangle
\quad
\text{ for $z<0$}\,,
\end{array}
\right.
\end{equation}
where the Moyal product $*$ is defined by
\begin{equation}
(f*g)(a,b)\equiv \lim_{a'\rightarrow a,\,b'\rightarrow b} e^{i\frac{\lambda}{4\pi}
(
\partial_{b} \partial_{a'}
-
\partial_a \partial_{b'}
)
}f(a,b) g(a',b')\,.
\end{equation}
This is the special case of the more general result
for an arbitrary gauge group that we now turn to.
\subsection{Non-Abelian gauge theories}
Here we consider a general $\mathcal N=2$ gauge theory with
arbitrary matter content.
Let us suppose that we have multiple line operators
$L_i\equiv L_{B_i,R_i}(\vec x=(0,0,z_i))$ located at various points
$\vec x=(0,0,z_i)$ on the 3-axis, ordered so that
\begin{equation}
z_1>z_2>\ldots > z_n\,.
\end{equation}
In the localization calculation, it suffices to consider
the Abelian configurations with magnetic charges $v_i$
associated with $B_i$ as only these
contribute to the path integral.
As is clear from the Maxwell case,
the holonomy at $z_i$ around $S^1$
is shifted by the magnetic fields $\propto v_j$ created by
$L_j$ for $j\neq i$:
\begin{equation}
a \rightarrow
a +\frac \lambda 2 \left(
\sum_{j<i} v_j-\sum_{j>i} v_j
\right)\,.
\end{equation}
Let us assume that the individual operator vevs are given by
\begin{equation}
\langle L\rangle
=
\sum_{v,w}
Z_{L,\text{total}}(a,b;v,w)
\equiv
\sum_{v,w}
e^{2\pi i(w\cdot a+v \cdot b)}
Z_L(a,m_f,\lambda;v,w)
\end{equation}
for some functions $Z_L(a,m_f,\lambda;v,w)$.
Then localization calculation yields
\begin{equation}
\langle L_1\cdot L_2\cdot \ldots \cdot L_n\rangle
=
\prod_{i=1}^n
\sum_{w_i}
\sum_{v_i}
Z_{L_i,\text{total}}
\left(a +\frac \lambda 2 \left(
\sum_{j<i} v_j-\sum_{j>i} v_j\right),
b\,
;v_i,w_i\right)\,,
\label{product-vev}
\end{equation}
One can easily see that (\ref{product-vev}) is the Moyal product
of the expectation values of individual operators
\begin{equation}
\langle L_1\cdot L_2\cdot \ldots \cdot L_n\rangle
= \langle L_1\rangle *\langle L_2\rangle * \ldots *
\langle L_n\rangle\,,
\end{equation}
where $*$ is defined by
\begin{equation}
(f*g)(a,b)\equiv
\left.
e^{i\frac{\lambda}{4\pi}
(
\partial_{b}\cdot \partial_{a'}
-
\partial_a\cdot \partial_{b'}
)
}f(a,b) g(a',b')
\right|_{a'=a,b'=b}
\end{equation}
with the natural product $\cdot$ between the derivatives inside the exponential.
As a concrete example, let us consider $SU(2)$ $\mathcal N=2^*$ theory.
We computed the vev of the charge-two 't Hooft operator in
(\ref{SU22starL20}).
As explained in \cite{Gomis:2011pf},
this operator corresponds to the product
of two minimal 't Hooft operators.
This is because the resolution of the singular
moduli space corresponds to separating the charge-two 't Hooft
operator into two minimal ones \cite{Kapustin:2006pk}.
Indeed one can check that the expression
(\ref{SU22starL20}) is precisely the Moyal product of
(\ref{SU22starL10}) with itself.
\subsection{Deformation quantization of the Hitchin moduli space}
\label{sec:Hitchin}
We are now going to explain that the noncommutative algebra structure
given by the Moyal multiplication above realizes a deformation quantization
of the Hitchin moduli space associated with the gauge theory.
In \cite{Gaiotto:2009we}, a correspondence between
certain $\mathcal N=2$ gauge theories and
punctured Riemann surfaces $C$ was discovered.
The correspondence is a main ingredient of the relation
\cite{AGT} between gauge theories and two-dimensional conformal field theories.
The correspondence is also manifested in the relation between
the gauge theories and the Hitchin systems on the Riemann surfaces.
This made it possible to study the integrable structure
\cite{Nekrasov:2009uh,2009AIPC.1134..154N,Nekrasov:2009ui}
as well as the low-energy dynamics
of these theories using the Hitchin system
on the Riemann surfaces \cite{Gaiotto:2009hg},
generalizing \cite{Donagi:1995cf}.
Let $A
=A_z dz+A_{\bar z} d\bar z
$ be a connection of a $G$-bundle over $C$,
and
$\varphi=\varphi_z dz+\bar\varphi_{\bar z} d\bar z$
an adjoint-valued 1-form.
They are assumed to possess prescribed singularities at the punctures.
The Hitchin moduli space is the space of solutions to
\begin{equation}
\begin{aligned}
&
\hspace{6mm}
F_{z\bar z}=[\varphi_z,\bar\varphi_{\bar z}]\,, \\
& D_{\bar z}\varphi_z=0\,,\quad
D_z \bar \varphi_{\bar z}=0\,,
\end{aligned}
\label{Hitchin-eq}
\end{equation}
up to $G$-gauge transformations.
The Hitchin moduli space is hyperK\"{a}hler ,
and therefore has a $\mathbb{CP}^1$ of complex structures $\mathcal J$,
each being a linear combination of three complex structures
$\mathcal J=I, J$, and $K$.
Each complex structure $\mathcal J$ is associated with a real symplectic form
$\omega_\mathcal J:=g \mathcal J$, as well as a holomorphic
symplectic form $\Omega_{\mathcal J}$.
For $\mathcal J=I, J, K$, these are given by
$\Omega_I=\omega_J+i\omega_K,
\Omega_J=\omega_K+i\omega_I,
\Omega_K=\omega_I+i\omega_J$.
In the original assignment of $I,J,K$ by Hitchin \cite{MR887284},
we are particularly interested in the complex structure $J$.
The combination $\mathcal A\equiv A+i\varphi$ is then holomorphic,
and (\ref{Hitchin-eq}) implies that\footnote{More precisely, the first of (\ref{Hitchin-eq}) combined with
the difference of the second and the third
is equivalent to the flatness of $\mathcal A$.
The $J$-holomorphic structure of the Hitchin moduli space
can be described by dropping the sum
of the second and the third equations,
and by taking the quotient with respect to $G_{\mathbb C}$
gauge transformations.
}
$\mathcal A$ is a flat $G_{\mathbb C}$ connection.
In terms of $\mathcal A$, $\Omega_J$ is given by
\begin{equation}
\Omega_J\propto\int_C {\rm Tr}\, \delta \mathcal A\wedge\delta \mathcal A\,.
\end{equation}
The $U(1)$ R-symmetry rotates the phases of $\varphi_z,
\bar\varphi_{\bar z}$, and $\Phi_0+i\Phi_9$,
and $\Omega_J$ transforms accordingly \cite{Gaiotto:2009hg}.
We focus on the one-punctured torus, which corresponds to
$SU(2)$ $\mathcal N=2^*$ theory.
Let us define generators of the first homology
so that the holonomy matrices $(A, B,M)$
along them satisfy the relation
\begin{equation}
AB=MBA\,.
\end{equation}
Here $M$ is the holonomy around a small circle surrounding the puncture,
and $A$ and $B$ are the holonomy matrices for the
usual A- and B-cycles.
Dehn's theorem \cite{dehn-breslau,MR956596}
allows us to label the non-self-intersecting closed curves
by two integers $(p,q)$ with equivalence
$(p,q)\sim (-p,-q)$.
They can be naturally identified with
the charges of
line operators in (\ref{charge-dict}) \cite{Drukker:2009tz}.
In particular, we have the correspondence
\begin{eqnarray}
\langle L_{0,1}\rangle &\leftrightarrow& {\rm Tr} A\,,\\
\langle L_{1,0}\rangle &\leftrightarrow& {\rm Tr} B\,,\\
\langle L_{1,\pm 1}\rangle &\leftrightarrow& {\rm Tr} A^{\pm 1}B\,.
\label{SU22star-corresp}
\end{eqnarray}
Let us consider the case $\lambda=0$.
From (\ref{SU22starL01}-\ref{SU22starL1pm1}) we find that
\begin{eqnarray}
\langle L_{0,1}\rangle_{\lambda=0}&=&
e^{2\pi i a}+e^{-2\pi i a}\,,
\label{2star01}
\\
\langle L_{1,0}\rangle_{\lambda=0}&=&
( e^{2\pi i b}+e^{-2\pi i b})
\left(
\frac{
\sin\left(2\pi a+\pi m\right)
\sin\left(2\pi a-\pi m\right)
}{
\sin^2\left(2\pi a\right)
}
\right)^{1/2}\,,
\label{2star10}
\\
\langle L_{1,\pm 1}\rangle_{\lambda=0}
&=&
( e^{2\pi i (b\pm a)}+e^{-2\pi i (b\pm a)})
\left(
\frac{
\sin\left(2\pi a+\pi m\right)
\sin\left(2\pi a-\pi m\right)
}{
\sin^2\left(2\pi a\right)
}
\right)^{1/2}\,.
\label{2star1pm1}
\end{eqnarray}
Replacing the arrows in (\ref{SU22star-corresp})
by equalities,
these expressions were exactly given as the definition of
the Darboux coordinates\footnote{In \cite{Nekrasov:2011bc} the Darboux coordinates were denoted by
$(\alpha, \beta)$, and are related to our $(a,b)$
by a trivial rescaling.
We also have ${\rm Tr} M=2\cos \pi m$.
} $(a,b)$ on the Hitchin moduli space with respect to
the symplectic structure $\Omega_J$!
Later in \cite{Dimofte:2011jd}, $(a,b)$
were identified with the complexification of the Fenchel-Nielsen
coordinates of Teichm\"uller space.
Here we see that both the coordinates $(a,b)$ and the symplectic
structure $\Omega_J$ arise naturally from the gauge theory on $S^1\times
\mathbb R^3$.
For $SU(2)$ $N_\text{F}=4$ theory, our gauge theory calculation
of the 't Hooft and dyonic operator vevs is not complete
due to the difficulty with monopole screening contributions.
The relation with Liouville theory and the formula
(\ref{Zmono-SU2NF4}) below suggests, however, that
$(a,b)$ are the complexified Fenchel-Nielsen coordinates
on the Hitchin moduli space
associated with the four-punctured sphere \cite{Cherkis:2000ft}.
\section{Gauge theory on $S^4$
and Liouville/Toda theories}
\label{sec:S4-CFT}
In this section we propose a precise relation between
the line operator vevs on $S^1\times \mathbb R^3$
and the corresponding difference operators
that act on the conformal blocks of Liouville and Toda
field theories.
We first motivate the correspondence by gauge theory considerations.
Then we will give an algorithm for computing
the line operator vevs on $S^1\times \mathbb R^3$
using two-dimensional CFT.
Let us consider the Liouville theory on
a genus $g$ Riemann surface with $n$ punctures
$C_{g,n}$.
The correlation function of primary fields $V_{\alpha_e}$
($e=1,\ldots, n$)
with momenta $\alpha_e$,
inserted at the punctures,
takes the form
\begin{equation}
\left \langle \prod V_{\alpha_e}
\right\rangle_{C_{g,n}}=
\int \left[\prod d\alpha_i \right]
\mathcal C(\alpha_i\,;\alpha_e)
|\mathcal F(\alpha_i;\alpha_e)|^2\,,
\label{CFT-corr}
\end{equation}
where the integral is over internal momenta $\alpha_i$
($i=1,\ldots, 3g-3+n$) and the function
$\mathcal C(\alpha_i\,;\alpha_e)$ is
a product of DOZZ three-point functions \cite{Dorn:1994xn,Zamolodchikov:1995aa}.
The conformal block $\mathcal F (\alpha_i;\alpha_e)$
depends on $\alpha_i, \alpha_e$, and the gluing parameters
$q_i$ holomorphically.
The central charge $c$ of Liouville theory is parametrized
as
\begin{equation}
c=1+6Q^2\,,
\quad\quad
Q={\mathsf b}+ {\mathsf b}^{-1}\,.
\end{equation}
The AGT correspondence \cite{AGT} states that
for ${\mathsf b}=1$ the correlation function (\ref{CFT-corr})
coincides with
the partition function of the corresponding $\mathcal N=2$
gauge theory on $S^4$ as defined by Pestun in \cite{Pestun:2007rz}.
The gluing parameters $q_i$ are related
to the complexified couplings
$\tau_i=\frac{\theta_i}{2\pi}+ \frac{4\pi i}{g_i^2}$
as $q_i=e^{2\pi i\tau_i}$.
Pestun's partition function contains as the north and south pole contributions
the Nekrasov instanton partition functions
defined in the Omega background \cite{Nekrasov:2002qd} .
The parameter ${\mathsf b}$ is related to
the equivariant parameters $\varepsilon_1, \varepsilon_2$
of the Omega background
as ${\mathsf b}^2= \varepsilon_1/\varepsilon_2$.
The path integral formulation
of the deformation $S^4_{\mathsf b}$ to ${\mathsf b}\neq 1$ is unknown at the time of writing.
Even for ${\mathsf b}\neq 1$,
it is expected that
(\ref{CFT-corr}) will be reproduced by
the partition function of the $\mathcal N=2$
gauge theory on $S^4$, deformed in
a certain way by a parameter $\mathsf b$.
For an $\mathcal N=2$ gauge theory with $SU(2)$ gauge groups
associated with a punctured Riemann surface $C_{g,n}$,
there is a correspondence between the charges $(B,R)$
of Wilson-'t Hooft operators and
a collection $\gamma$ of non-self-intersecting
closed curves on $C_{g,n}$ \cite{Drukker:2009tz}.
In \cite{Alday:2009fs,Drukker:2009id}
it was shown that there exists a difference operator $\Lambda_\gamma$,
the Verlinde operator, whose action
on $\mathcal F(\alpha_i;\alpha_e)$ we denote by\footnote{For the Verlinde operator in Liouville theory,
corresponding to a connected closed curve on the Riemann surface,
our normalization of the operator
agrees with \cite{Alday:2009fs}.
Our operator is $2\cos {\mathsf b} Q$ times those in \cite{Drukker:2009id,Gomis:2010kv}.
}
\begin{equation}
\mathcal F(\alpha_i;\alpha_e)
\rightarrow
[\Lambda_\gamma\cdot \mathcal F]
(\alpha_i;\alpha_e)
\,.
\end{equation}
The same papers demonstrated that, for $\mathsf b=1$, the expectation value of the
Verlinde operator defined as
\begin{equation}
\int \left[\prod d\alpha_i \right]
\mathcal C(\alpha_i\,;\alpha_e)
\overline{\mathcal F(\alpha_i;\alpha_e)}
[\Lambda_\gamma\cdot \mathcal F]
(\alpha_i;\alpha_e)
\label{Verlinde-vev-F}
\end{equation}
reproduces the expectation value of the Wilson loop on $S^4$
computed by Pestun \cite{Pestun:2007rz}.
The agreement of (\ref{Verlinde-vev-F}) with the 't Hooft loop
expectation value, again for $\mathsf b=1$, was
more recently verified in \cite{Gomis:2011pf}.
This was done by performing a localization calculation for 't Hooft loops
placed along a large circle, called the equator, of $S^4$.
The neighborhood of the equator is approximately $S^1\times \mathbb R^3$,
therefore much of the analysis overlaps the present paper.
Because of the curvature, however,
in the orthonormal frame of the metric such that
the conformal Killing spinor is periodic,
the hypermultiplet becomes antiperiodic.
We expect that this property persists for ${\mathsf b}\neq 1$.
The effect of antiperiodicity is to multiply
the right hand side of (\ref{ind-Dhm}) by an extra factor $e^{ i \varepsilon/2}$.
Thus we conjecture, and the AGT correspondence suggests,
that the 't Hooft loop vev on $S^4$
for general $\mathsf b$ is given by
\begin{equation}
\begin{aligned}
\langle T_B\rangle_{S^4}&
= \int_{i \mathfrak t} da
\sum_v
Z_\text{pole}(a+{\mathsf b}^2v/2,\bar q)
Z_\text{equator}(a;B,v)
Z_\text{north}(a-{\mathsf b}^2v/2,q)
\\
&= \int_{i \mathfrak t} da\,
Z_\text{pole}(a,\bar q)
\sum_v
e^{-(\mathsf b^2/2) v\cdot \partial_a}
Z_\text{equator}(a;B,v)
e^{-(\mathsf b^2/2) v\cdot \partial_a}
Z_\text{north}(a,q)
\,,
\end{aligned}
\label{thooft-S4}
\end{equation}
where
\begin{equation}
Z_\text{pole}(a,q)=e^{-S_\text{cl}(a)}
Z_\text{1-loop}(a)
Z_\text{inst}(a,q)
\label{poles}
\end{equation}
is the Nekrasov partition function
including the classical, one-loop, and instanton contributions
on $\mathbb C^2$,\footnote{Since we are primarily interested in
the equator contributions we suppress the dependence
on
$\varepsilon_1$, $\varepsilon_2$, $\mathsf m_f$, and $\lambda$
in (\ref{poles}).
The mass parameters $\mathsf m_f$
as well as the Coulomb moduli $a$ are pure imaginary.
}
and the equator contribution
\begin{equation}
Z_\text{equator}(a;B,v)
= Z_\text{1-loop}(a,{\mathsf m}_f+1/2,{\mathsf b}^2;v)
Z_\text{mono}(a,{\mathsf m}_f+1/2,{\mathsf b}^2;B,v)\,.
\label{eq-cont}
\end{equation}
is given in terms of the one-loop determinant in (\ref{1-loop-total})
and the monopole screening contribution in (\ref{mono-total}).
The shift in mass is due to the antiperiodicity of hypermultiplets
mentioned above.
For ${\mathsf b}=1$ (\ref{thooft-S4}) was established in \cite{Gomis:2011pf},
where the definitions of $Z_\text{1-loop}$ and $Z_\text{mono}$
were slightly different due to the the shift in the mass.
The second equality in (\ref{thooft-S4}) involves
a shift of integration contours
and integration by parts.
For some examples in Liouville theory,
it was checked that the shift of contours
does not encounter poles \cite{Gomis:2011pf}.
Based on the conjectured relation (\ref{thooft-S4}) between the line operator vevs on $S^4$
and $S^1\times \mathbb R^3$
for general $\lambda={\mathsf b}^2$,
we propose that
the vev in the theory on $S^1\times \mathbb R^3$ can be
obtained from the Verlinde operator in Liouville theory
by the following algorithm.
This algorithm was used in \cite{Gomis:2011pf} in the case ${\mathsf b}=1$
to read off $Z_\text{equator}$ from the Verlinde operator.
First we change the normalization of the conformal block and define\footnote{The Verlinde operator in \cite{Alday:2009fs}
was computed in the standard normalization for conformal blocks
\cite{Belavin:1984vu}.
Yet another normalization introduced in \cite{Ponsot:1999uf} was
used to calculate the Verlinde operators in \cite{Drukker:2009id},
and in this basis the operators are free of square-roots \cite{Dimofte:2011jd}.
}
\begin{equation}
\mathcal B(\alpha_i;\alpha_e)\equiv \mathcal C(\alpha_i;\alpha_e)^{1/2}
\mathcal F(\alpha_i;\alpha_e)
\label{B-block}
\end{equation}
using the square root of the function $\mathcal C(\alpha_i;\alpha_e)$
that appears in the correlation function (\ref{CFT-corr}).
With the one-loop factor in (\ref{poles}) whose precise definition
was given in \cite{Gomis:2011pf},
we expect that
$\mathcal B(\alpha)=Z_\text{pole}(a)$
with the identification $\alpha_i=Q/2+a_i/{\mathsf b}$.
In this normalization, the Liouville correlation function is
simply given by
\begin{equation}
\left \langle \prod V_{\alpha_e}
\right\rangle_{C_{g,n}}=
\int \left[\prod d\alpha_i \right]
\,
|\mathcal B(\alpha_i;\alpha_e)|^2\,,
\label{CFT-corr2}
\end{equation}
where we used the fact that in the physical range of
Liouville momenta, the function $\mathcal C(\alpha_i;\alpha_e)$ is real.
The Verlinde operator acts on $\mathcal B(\alpha)$ as
the difference operator defined by
\begin{equation}
[\mathcal L_\gamma\cdot \mathcal B]
(\alpha_i;\alpha_e)
\equiv
\mathcal C(\alpha_i;\alpha_e)^{1/2}
[\Lambda_\gamma\cdot \mathcal F](\alpha_i;\alpha_e)\,.
\label{Ver-on-B}
\end{equation}
Its vev is then given by
\begin{equation}
\int \left[\prod d\alpha_i \right]
\overline{\mathcal B(\alpha_i;\alpha_e)}
[\mathcal L_\gamma\cdot \mathcal B]
(\alpha_i;\alpha_e)\,.
\label{Verlinde-vev-B}
\end{equation}
The operator algebra of $\mathcal L_\gamma$
is isomorphic to that of $\Lambda_\gamma$.
In the case $\gamma$ is purely magnetic,
we conjecture that $\mathcal L_\gamma$ is related to the 't Hooft loop
$T_B$ above as
\begin{equation}
\mathcal L_\gamma= \sum_v
e^{-(\mathsf b^2/2) v\cdot \partial_a}
Z_\text{equator}(a;B,v)
e^{-(\mathsf b^2/2) v\cdot \partial_a}
\label{ver-general}
\end{equation}
up to an overall constant.
For more general dyonic charges
the Verlinde operator takes the form
\begin{equation}
\mathcal L
=\sum_{v,w}
e^{-(\mathsf b^2/2) v\cdot \partial_a}
e^{2\pi i w\cdot a}
Z_L\left(a,{\mathsf m}_f+\frac 1 2, {\mathsf b}^2;v,w\right)
e^{-(\mathsf b^2/2) v\cdot \partial_a}
\,,
\label{VZZ}
\end{equation}
with some functions $Z_L(a,m_f, \lambda;v,w)$.
For a product of $SU(2)$'s, our Lie algebra convention is such that
$v\cdot \partial_a=\sum_i v_i\frac{\partial}{\partial a_i}$ and
$w\cdot a =\sum_i{\rm Tr}\left[
\text{diag}(w_i,-w_i)
\text{diag}(a_i,-a_i)
\right]=2 \sum_i w_i a_i$,
with $v_i$ and $w_i$ being half-integers.
The ``highest'' $v$ (corresponding to the highest weight of the Langlands-dual
representation) and $w$ have $v_i=p_i/2$ and $w_i=q_i/2$,
where $(p_i,q_i)$ are the Dehn-Thurston
parameters \cite{Drukker:2009tz}.
We conjecture that the line operator vevs are
given in terms of these functions as
\begin{equation}
\hspace{-2mm}
\langle L\rangle_{S^4}=
\hspace{-1mm}
\int_{i \mathfrak t} da\,
Z_\text{pole}(a,\bar q)
\sum_{v,w}
e^{-\frac{\mathsf b^2}{2} v\cdot \partial_a}
e^{2\pi i w\cdot a}
Z_L\left(a,{\mathsf m}_f+\frac 1 2, {\mathsf b}^2;v,w\right)
e^{-\frac{\mathsf b^2}{2} v\cdot \partial_a}
Z_\text{pole}(a,q)
\label{L-S4-conj}
\end{equation}
on $S^4$ and
\begin{equation}
\langle L\rangle_{S^1\times \mathbb R^3}=
\sum_{v,w}
e^{2\pi i( w\cdot a+v\cdot b)}
Z_L(a, m_f, \lambda;v,w)
\label{L-S1R3-conj}
\end{equation}
on $S^1\times \mathbb R^3$.
We have focused so far on the correspondence \cite{AGT} between
the gauge theories whose gauge group is a product of $SU(2)$'s
and Liouville theory on the corresponding Riemann surface,
but we also propose that the relation (\ref{thooft-S4})
should hold for more general gauge groups and Toda theories \cite{AGT,Wyllard:2009hg}.
Some examples of Verlinde operators in Toda theories were computed in
\cite{Passerini:2010pr,Gomis:2010kv}.
We conjecture that the Verlinde operators
in Toda theories
are precisely
related to the line operator vevs on $S^4$ and $S^1\times\mathbb R^3$
via the equations (\ref{VZZ}), (\ref{L-S4-conj}), and
(\ref{L-S1R3-conj}).
We observe that the Verlinde operator (\ref{VZZ})
is related to the vev
(\ref{L-S1R3-conj})
on $S^1\times \mathbb R^3$
precisely by the Weyl transform (ordering)\footnote{For a 2-dimensional phase space parametrized by $(q,p)$,
the operator $\mathcal O$ and its inverse Weyl transform $f$
are
related by
$
f(q,p)
=\int d\sigma e^{-\frac{i}\hbar p \sigma}
\langle q
|
e^{\frac{i} {2\hbar} \sigma \hat p}
\mathcal O(\hat q,\hat p)
e^{\frac{i} {2\hbar} \sigma \hat p}
|
q
\rangle
\,,
\mathcal O
=
\frac{1}{(2\pi)^2 \hbar}
\int d\sigma d\tau dq dp
e^{-i\tau(\hat q-q)-\frac{i}\hbar\sigma(\hat p-p)}
f(q,p)\,,
$
where $[\hat q,\hat p]=i\hbar$, $\hat q|q\rangle=q|q\rangle$,
$\langle q|q'\rangle=\delta(q-q')$ \cite{Harvey:2001yn}.
}
\begin{equation}
\langle L\rangle_{S^1\times \mathbb R^3}
\quad
\stackrel{\text{Weyl}}{\Longrightarrow}
\quad
\mathcal L\,.
\end{equation}
The parameter $-b$ plays the role of the canonical momentum:
\begin{equation}
\begin{aligned}
b&\leftrightarrow i\frac{\lambda}{2\pi}\frac{\partial}{\partial a}
\hspace{11mm}
\text{in general,}
\\
b_i&\leftrightarrow i\frac{\lambda}{4\pi}\frac{\partial}{\partial a_i}
\hspace{10mm}
\text{for $SU(2)$ and Liouville. }
\end{aligned}
\end{equation}
Thus our proposal (\ref{thooft-S4}) implies that the Verlinde operators
are the Weyl transform of the line operator vevs on $S^1\times \mathbb R^3$,
when the gauge theory has a Lagrangian description.
It is very natural to conjecture that this relation
should hold even when the gauge theory does not admit a Lagrangian
description \cite{Argyres:2007cn,Gaiotto:2009we}.
The mass shift in (\ref{eq-cont}) and (\ref{ver-general})
is consistent with a somewhat confusing aspect of the correspondence
\cite{Gaiotto:2009we}
between $\mathcal N=2$ gauge theories and Riemann surfaces.
Namely the massless limit of a gauge theory
corresponds to removing a puncture in the Hitchin system
\cite{Gaiotto:2009hg,Gaiotto:2010be},
while it corresponds to tuning external momenta
to special values in Liouville/Toda theories \cite{AGT,Wyllard:2009hg,Okuda:2010ke} keeping the puncture.
In gauge theories the shift arises due to the difference
in the geometries where the theories live.
Below we demonstrate our proposal with several examples.
\subsection{$SU(2)$ $\mathcal N=2^*$}
This theory corresponds to the Liouville theory on the one-punctured torus
\cite{AGT}.
Let $C(\alpha_1,\alpha_2,\alpha_3)$ be the DOZZ three-point function
of Liouville theory \cite{Dorn:1994xn,Zamolodchikov:1995aa}.
We denote the internal and external Liouville momenta
by $\alpha$ and $\alpha_e$ respectively.
The Verlinde loop operator that corresponds to the minimal
't Hooft operators acts on the conformal block
as
\begin{equation}
[ \mathcal L_{1,0}\cdot \mathcal F](\alpha,\alpha_e)=
\sum_\pm
H_\pm(\alpha)
\mathcal F\left(\alpha\pm {\mathsf b}/{2}\,,\alpha_e\right)\,.
\end{equation}
This implies the following expression, conjectural for $\mathsf b\neq 1$,
of the minimal 't Hooft operator vev in the $\mathcal{N}=2^*$ theory on $S^4$:
\begin{align}
\langle\,
L_{1,0}\,
\rangle_{S^4}
=
\int_{Q/2+i\mathbb R} d\alpha\,
C(\alpha,\alpha_e,Q-\alpha)
\sum_\pm
\overline{\mathcal F (\alpha,\alpha_e)}
H_\pm(\alpha)
\mathcal F\left(\alpha\pm {\mathsf b}/{2}\,,\alpha_e\right)\,.
\label{tHooftLiou2star}
\end{align}
The map between the Liouville and gauge theory parameters is given by
\begin{align}
\alpha=\frac{Q}{2}+ \frac{a}{{\mathsf b}},\quad \alpha_e=\frac{Q}{2}+ \frac{{\mathsf m}}{{\mathsf b}}.
\end{align}
The coefficients $H_\pm$ are known to be \cite{Alday:2009fs}
\begin{align}
H_\pm(\alpha)
=\frac{\Gamma(\pm 2 a)\Gamma(\pm 2 a+{\mathsf b} Q)}
{\Gamma(\pm 2 a + {\mathsf m}+{\mathsf b} Q/2)\Gamma(\pm 2 a-{\mathsf m}+{\mathsf b} Q/2)}.
\end{align}
By performing the manipulations explained above,
in Appendix \ref{app:Liou-2star} we obtain
\begin{align}
\mathcal L_{1,0}
=
\sum_\pm e^{\pm \frac{1}{4} {\mathsf b}^2\partial_a}
\left(
\prod_\pm
\frac{
{ \cos ( 2\pi a \pm \pi {\mathsf m})
}}
{
\sin( 2\pi a \pm \frac{\pi}{2}{\mathsf b}^2)
}
\right)^{1/2}
e^{\pm \frac{1}{4} {\mathsf b}^2\partial_a}\,.
\label{2star-L10}
\end{align}
This is indeed related to
the 't Hooft operator vev
(\ref{SU22starL10})
by the Weyl transform above.
\subsection{$SU(N)$ $\mathcal N=2^*$
}
The Verlinde operator corresponding to the 't Hooft operator
with charge $B=(1,0^{N-1})$, acting on the Toda conformal block
for the one-punctured torus, was computed in
\cite{Gomis:2010kv} in the standard normalization.
In Appendix \ref{app:Toda-2star} we convert it
to a difference operator acting on the block $\mathcal B$
in the normalization that absorbs the square root of
the three-point function.
We find
\begin{equation}
\mathcal L_{B=(1,0^{N-1})}=
\sum_{l=1}^N
e^{-\frac{\mathsf b^2}{2} h_l\cdot \partial_a}
\left(
\prod_\pm
\prod_{j\neq l}
\frac{
\cos\pi (a_{lj}\pm \mathsf m)
}{
\sin\pi (a_{lj}\pm\lambda/2)
}
\right)
e^{-\frac{\mathsf b^2}{2} h_l\cdot \partial_a}
\,,
\label{Ver-SUN-star-thooft}
\end{equation}
where $(h_l)_j=\delta_{jl}-1/N$.
Note that $h_l$ are the coweights that correspond to
the weights in the fundamental representation
of the Langlands dual group.
The Verlinde operator (\ref{Ver-SUN-star-thooft})
is the Weyl transform of the vev
(\ref{UN2starBpm10N-1}) on $S^1\times\mathbb R^3$ as expected.
\subsection{$SU(2)$ $N_\text{F}=4$}
To compare with gauge theory calculations,
we relate $a$ and ${\mathsf m}_f$ to $\alpha$ and $\alpha_e$ by
\begin{equation}
\begin{aligned}
& \alpha=\frac{Q}{2} + \frac{a}{{\mathsf b}}\,,
\quad
\alpha_1=\frac{Q}{2} + \frac{{\mathsf m}_1-{\mathsf m}_2}{2{\mathsf b}}\,,
\quad
\alpha_2=\frac{Q}{2} + \frac{{\mathsf m}_1+{\mathsf m}_2}{2{\mathsf b}}\,,
\\
&
\hspace{15mm}
\alpha_3=\frac{Q}{2} + \frac{{\mathsf m}_3+{\mathsf m}_4}{2{\mathsf b}}\,,
\quad
\alpha_4=\frac{Q}{2} + \frac{{\mathsf m}_3-{\mathsf m}_4}{2{\mathsf b}}\,.
\end{aligned}
\end{equation}
For the minimal Wilson operator,
the corresponding Verlinde operator is
\begin{equation}
\mathcal L_{0,1}=e^{2\pi i a}+e^{-2\pi i a}\,.
\end{equation}
In Appendix (\ref{app:Toda-NF=2N}), we show that
\begin{equation}
\begin{aligned}
\mathcal L_{2,0}
&=
\sum_\pm e^{\pm \frac{1}{2} {\mathsf b}^2\partial_a}
\left(
\frac{
\prod_{f=1}^4\prod_{s=\pm}
\cos (\pi a +s \pi {\mathsf m}_f)
}{
\sin(2\pi a + \pi {\mathsf b}^2)
\sin^2 (2\pi a )
\sin( 2\pi a- \pi {\mathsf b}^2)
}
\right)^{1/2}
e^{\pm \frac{1}{2} {\mathsf b}^2\partial_a}
\\
&\quad
\quad\quad
-\frac 1 2
\cos\pi( {\mathsf b}^2-\sum_f {\mathsf m}_f)
+
\sum_{s=\pm} \frac{
\prod_{f=1}^4 \cos \pi \left(-{\mathsf b}^2/2 +s a + {\mathsf m}_f\right)
}{
\sin( 2\pi a) \sin \pi (s {\mathsf b}^2 -2 a)
}\,.
\label{Ver20}
\end{aligned}
\end{equation}
In view of (\ref{VZZ}) and (\ref{L-S1R3-conj}),
this is related to
the 't Hooft operator vev (\ref{U2NF4})
in the $U(2)$ theory with $a_1=-a_2=a$
by the Weyl transform,
up to an $a$-independent term $-\frac12\cos\pi({\mathsf b}^2-\sum {\mathsf m}_f)$.
The whole expression is invariant under $a\rightarrow -a$
as well as under the action of the $SO(8)$ Weyl group.\footnote{Four generators of the Weyl group of the $SO(8)$ flavor group
act on the masses
as ${\mathsf m}_1\leftrightarrow {\mathsf m}_2$,
${\mathsf m}_2\leftrightarrow {\mathsf m}_3$,
${\mathsf m}_3\leftrightarrow {\mathsf m}_4$,
and
${\mathsf m}_3\leftrightarrow - {\mathsf m}_4$ respectively.
Agreement up to an additive constant
is almost as much as one can hope for.
Without $SO(8)$ Weyl invariance, however,
(\ref{U2NF4}) with $a_1=a=-a_2$
cannot be the answer for $SU(2)$ gauge theory.
}
We see that Liouville theory in fact fixes the $a$-independent
additive constant.
This constant is such that the whole second line
of (\ref{Ver20}) vanishes in the limit $a\rightarrow \pm i\infty$.
By the argument given above, then,
Liouville theory predicts that the $SU(2)$ theory with $N_\text{F}=4$
has
\begin{equation}
Z_\text{mono}(a,m_f;2,0)=
-\frac{1}{2}\cos\pi( \lambda-\sum_f m_f)
-\sum_{s=\pm} \frac{
\prod_{f=1}^4 \sin \pi \left (s a - m_f+\lambda/2\right)
}{
\sin( 2\pi a) \sin \pi (s \lambda +2 a)
}\,.
\label{Zmono-SU2NF4}
\end{equation}
Thus the minimal 't Hooft operator vev
on $S^1\times \mathbb R^3$
should be%
\footnote{%
Essentially the same expression has been obtained
purely from quantization of the Hitchin system \cite{Teschner-KITP}.
}
\begin{equation}
\begin{aligned}
\langle L_{2,0}\rangle
=&
(e^{4\pi i b}+e^{-4\pi i b})
\left(
\frac{
\prod_\pm \prod_{f=1}^4 \sin\pi(a\pm m_f)}
{
\sin^22\pi a \prod_\pm \sin\pi(2a\pm\lambda)
}
\right)^{1/2}
\\
&
\quad\quad
-\frac{1}{2}\cos\pi( \lambda-\sum_f m_f)
-\sum_{s=\pm} \frac{
\prod_{f=1}^4 \sin \pi \left (s a - m_f+\lambda/2\right)
}{
\sin( 2\pi a) \sin \pi (s \lambda +2 a)
}\,.
\end{aligned}
\label{minimal-thooft-from-liouville}
\end{equation}
In Appendix \ref{sec:cl-holo}, we show that
the $\lambda=0$ limit of this expression coincides with
the classical holonomy on the four-punctured sphere
written in terms of the complexified Fenchel-Nielsen coordinates.
\subsection{$SU(N)$ $N_\text{F}=2N$}
The Verlinde operator in the $SU(N)$ superconformal QCD
corresponding to the 't Hooft operator
with $B=(1,-1,0^{N-2})$ was computed in
\cite{Gomis:2010kv}.
This was done in the standard normalization,
and in Appendix \ref{app:Toda-NF=2N}, we convert the operator to
the difference operator acting on the block $\mathcal B$.
It is given up to a multiplicative constant by
\begin{eqnarray}
\hspace{-10mm}
&&
\hspace{6mm} \mathcal L_{B=(1,-1,0^{N-2})}
\nonumber\\
\hspace{-10mm}
&&=
\hspace{-2mm}
\mathop{\sum_{1\leq j,k\leq N}}_{j\neq k}
\hspace{-2mm}
e^{-\frac{{\mathsf b}^2}{4}
e_{jk}
\cdot\partial_a}
\frac{
\left[\prod_{f=1}^N
\cos \pi (a_j-{\mathsf m}_f)
\cos\pi (a_k-{\mathsf m}_f)
\right]^{\frac 1 2}
}{
\sin\pi a_{jk}
{\displaystyle \prod_\pm }
\bigg[
\sin \pi (a_{jk}\pm{\mathsf b}^2)
{\displaystyle
\prod_{i\neq j,k}}
\sin\pi(a_{ji}\pm \frac{{\mathsf b}^2}{2})
\sin\pi(a_{ik}\pm \frac{{\mathsf b}^2}{2})
\bigg]^{\frac 1 2}
}
e^{-\frac{{\mathsf b}^2}{4} e_{jk}
\cdot\partial_a}
\nonumber\\
\hspace{-10mm}&&
\quad\quad\quad\quad
+\sum_{k=1}^N
\frac{
\prod_{f=1}^{2N}
\cos \pi
\left(
a_k-{\mathsf m}_f +\frac{{\mathsf b}^2}{2}
\right)
}
{
\prod_{i\neq k}
\sin \pi a_{ki}
\sin \pi\left(- a_{ki}-{\mathsf b}^2\right)
}
\\
\hspace{-10mm}&&
\quad\quad\quad\quad
+(-1)^{N-1}
\frac{e^{N \pi i (\sum_{f>N} \mathsf m_f-\sum_{f\leq N} \mathsf m_f)/N}
\sin\pi \mathsf b^2}
{\sin(\pi(N-2)\mathsf b^2)}
-\frac 1 2 \cos \pi (\mathsf b^2 -\sum_f \mathsf m_f)
\,.
\nonumber
\end{eqnarray}
Here $e_{jk}\equiv h_j-h_k$ are the coroots.
This implies that for the minimal 't Hooft operator
in the $SU(N)$ theory with $N_\text{F}=2N$,
the vev on $S^1\times\mathbb R^3$ is given by
\begin{eqnarray}
&& \langle T_{B=(1,-1,0^{N-2})}\rangle
\nonumber \\
&=&
\mathop{\sum_{1\leq j,k\leq N}}_{j\neq k}
e^{\pi i(b_j-b_k)}
\frac{
\left[
\prod_{f=1}^N
\sin\pi (a_j-m_f)
\sin\pi (a_k-m_f)
\right]^{1/2}
}{
\sin\pi a_{jk}
{\displaystyle \prod_\pm }
\left[
\sin \pi (a_{jk}\pm\lambda)
{\displaystyle
\prod_{i\neq j,k}}
\sin\pi(a_{ji}\pm \lambda/2)
\sin\pi(a_{ik}\pm \lambda/2)
\right]
^{1/2}
}
\nonumber
\\
&&
\quad\quad\quad\quad
+\sum_{k=1}^N
\frac{
\prod_{f=1}^{2N}
\sin \pi
\left(
a_k- m_f +\lambda/2
\right)
}
{
\prod_{i\neq k}
\sin \pi a_{ki}
\sin \pi\left(- a_{ki}-\lambda\right)
}
\label{SUN-SQCD}
\\
&&
\quad\quad\quad\quad
+(-1)^{N-1}
\frac{e^{\pi i (\sum_{f>N} m_f-\sum_{f\leq N} m_f)}
\sin\pi \lambda}
{\sin(\pi(N-2)\lambda )}
+\frac{(-1)^{N-1}} 2 \cos \pi (\lambda -\sum_f m_f)
\,.\nonumber
\end{eqnarray}
This is identical to the $U(N)$ result (\ref{UNSQCD}) up to the terms
independent of $a$ and $b$.
The expression (\ref{SUN-SQCD}) is a prediction of Toda theory
for the $SU(N)$ gauge theory.
\section{Discussion}
\label{sec:discussion}
Let us conclude with remarks on future directions and related works.
We focused on conformal $\mathcal N=2$ gauge theories because localization
calculations are the cleanest for them.
Line operators in non-conformal asymptotically free theories also exhibit
rich dynamics \cite{Gaiotto:2010be} and the spectrum of BPS states is often simpler.
The easiest way to compute correlation functions in such theories would be
to start with a conformal theory and decouple some matter fields
by sending their mass to infinity.
It would be interesting to study this limit in detail.
Our calculation of the line operator vevs, or the supersymmetric index (\ref{trace-def}),
in terms of the complexified Fenchel-Nielsen coordinates made use of
the equivariant index theorem.
It is amusing to note that the calculation of the supersymmetric index
in terms of the Fock-Goncharov coordinates can also be formulated
in terms of an index theorem,
but applied to the moduli space constructed from the Seiberg-Witten
prepotential governing the IR dynamics \cite{Lee:2011ph,Kim:2011sc} .
In our computational scheme for the monopole screening contributions $Z_\text{mono}(B,v)$
in $\mathcal N=2^*$ theory,
the 't Hooft operator is S-dual to the Wilson operator
in a product of fundamental representations.
As such the 't Hooft operator is reducible, i.e., it can be written as a
linear combination of other line operators with positive coefficients.
Related to this is the fact that the 't Hooft operator vev (\ref{U2NF4}) in the $G=U(2)$ theory
with $N_\text{F}=4$ fundamental hypermultiplets
becomes $SO(8)_\text{F}$ Weyl-invariant not just by substituting
$(a_1,a_2)\rightarrow (a,-a)$,
but only after adding an $a$-independent term in (\ref{Ver20}).
It is important to develop a method intrinsic to irreducible
line operators for gauge group $SU(N)$ rather than $U(N)$.
This may involve decomposing the cohomology
of the monopole moduli space into irreducible representations
of the Langlands dual gauge group \cite{Witten:2009mh}
and incorporate the computation of operator product expansions
\cite{Kapustin:2006pk,Kapustin:2007wm,Saulina:2011qr}.
We found that the line operators in $\mathcal N=2$ theories
on $S^1\times \mathbb R^3$
realize a deformation quantization of the Hitchin moduli space.
We expect that this can be explained in the framework of
\cite{Nekrasov:2010ka}, by dimensionally reducing the theory on the
circle parametrized by $\tau$ as well as the one
parametrized by the polar angle in the 12-plane.
We would obtain a $(4,4)$ sigma model on a half plane whose target space
is the Hitchin moduli space, and the
boundary condition would correspond to
the canonical coisotropic brane.
Liouville/Toda conformal blocks arise as open string states
by including another boundary mapped to the brane of opers.
It would be interesting to study these systems in more detail and
understand the appearance of the Weyl transform.
Some of the results in \cite{Gaiotto:2009hg}
obtained by the wall-crossing formula
can be reproduced from our results that are obtained
directly by localization calculations.
It would be interesting to further explore the relation between
the UV and IR theories as well as the integrability aspects of the
line operators.
\section*{Acknowledgements}
We are grateful to Nadav Drukker, Jaume Gomis, Anton Kapustin,
Yoichi Kazama, Satoshi Nawata, Andy Neitzke,
Vasily Pestun,
Natalia Saulina, and Masahito Yamazaki for fruitful discussions.
T.O. and M.T. also thank the KITP for providing us with
a stimulating environment where part of this work was done.
The research at KITP was supported in part by DARPA under Grant No.
HR0011-09-1-0015 and by National Science Foundation under Grant
No. PHY05-51164.
T.O. thanks the Perimeter Institute for warm hospitality.
The research of Y.I is supported in part by
a JSPS Research Fellowship for Young Scientists.
The research of T.O. is supported in part
by Grant-in-Aid for Young Scientists (B) No. 23740168,
Grant-in-Aid for Scientific Research (B) No. 20340048,
and Institutional Program for Young Researcher Overseas Visits
No. R10 from the Japan Society for the Promotion of Science.
The research of M.T. is supported in part
by JSPS Grant-in-Aid for Creative Scientific Research No. 19GS0219.
\vfill\eject
|
\section{Supplementary Material}
This Supplementary Material contains detailed derivations of some of
the results presented in the main text. Section~\ref{czzb} derives the
classical Ziv-Zakai bounds, Sec.~\ref{qzzb_gauss} calculates a quantum
Ziv-Zakai bound (QZZB) when the photon-number distribution can be
approximated as continuous and Gaussian, Sec.~\ref{qzzb_coh}
calculates the bound for coherent states, and Sec.~\ref{qzzb_rect}
calculates the bound for rectangle states.
\subsection{\label{czzb}Derivation of the classical Ziv-Zakai bounds}
Let
\begin{align}
\sigma &\equiv |\tilde X(Y)-X|
\end{align}
be a nonnegative random variable. The mean-square error becomes
\begin{align}
\Sigma &= \int_0^\infty ds P_\sigma(s)s^2,
\end{align}
where $P_\sigma(s)$ is the probability density of $\sigma$. With
\begin{align}
P_\sigma(s) &= -\diff{}{s}\op{Pr}\bk{\sigma \ge s}
\end{align}
and integration by parts, $\Sigma$ becomes
\begin{align}
\Sigma &= 2\int_0^\infty ds s\op{Pr}\bk{\sigma \ge s}
= \frac{1}{2}\int_0^\infty d\tau \tau\op{Pr}\bk{|\tilde X-X|\ge \frac{\tau}{2}}.
\end{align}
Since both $\tau$ and $\op{Pr}(|\tilde X-X|\ge \tau/2)$ are
nonnegative, one can find a lower bound on $\Sigma$ by bounding
$\op{Pr}(|\tilde X-X|\ge \tau/2)$. Rewrite $\op{Pr}(|\tilde X-X|\ge
\tau/2)$ as follows:
\begin{align}
\op{Pr}\bk{|\tilde X-X|\ge \frac{\tau}{2}}
&= \op{Pr}\bk{\tilde X-X> \frac{\tau}{2}}
+\op{Pr}\bk{\tilde X-X \le -\frac{\tau}{2}}
\\
&= \int_{-\infty}^{\infty} dx_0 P_X(x_0)\op{Pr}\bk{\tilde X-X> \frac{\tau}{2}\bigg|X = x_0}
\nonumber\\&\quad
+
\int_{-\infty}^{\infty} dx_1 P_X(x_1)\op{Pr}\bk{\tilde X-X\le -\frac{\tau}{2}\bigg|X = x_1},
\end{align}
and let
\begin{align}
x_0 &= x, & x_1 &= x+\tau.
\end{align}
This yields
\begin{align}
\op{Pr}\bk{|\tilde X-X|\ge \frac{\tau}{2}}
&= \int_{-\infty}^{\infty} dx
\bigg[P_X(x)\op{Pr}\bk{\tilde X> x+\frac{\tau}{2}\bigg|X = x}
\nonumber\\&\quad
+P_X(x+\tau)\op{Pr}\bk{\tilde X\le x+\frac{\tau}{2}\bigg|X = x+\tau}\bigg]
\label{eq1}
\\
&= \int_{-\infty}^{\infty} dx
[P_X(x)+P_X(x+\tau)]\nonumber\\&\quad\times
\BK{P_0\op{Pr}\bk{\tilde X> x+\frac{\tau}{2}\bigg|X = x}
+
P_1\op{Pr}\bk{\tilde X\le x+\frac{\tau}{2}\bigg|X = x+\tau}},
\label{eq2}
\end{align}
where
\begin{align}
P_0 &\equiv \frac{P_X(x)}{P_X(x)+P_X(x+\tau)},
&
P_1 &\equiv \frac{P_X(x+\tau)}{P_X(x)+P_X(x+\tau)} = 1-P_0.
\end{align}
Now consider a binary hypothesis testing problem with hypotheses
\begin{align}
\mathcal H_0: X &= x, & \mathcal H_1: X &= x+\tau,
\end{align}
prior probabilities
\begin{align}
\op{Pr}(\mathcal H_0) &= P_0,
&\op{Pr}(\mathcal H_1) &= P_1,
\end{align}
observation probability densities
\begin{align}
P_Y(y|\mathcal H_0) &= P_{Y|X}(y|x),
&
P_Y(y|\mathcal H_1) &= P_{Y|X}(y|x+\tau),
\end{align}
and the following suboptimal decision rule:
\begin{align}
\op{choose }\mathcal H_0 \op{ if }\tilde X &\le x + \frac{\tau}{2},
&
\op{choose }\mathcal H_1 \op{ if }\tilde X &> x + \frac{\tau}{2}.
\end{align}
The error probability of this hypothesis testing problem is then
precisely given by the expression in the curly brackets in
Eq.~(\ref{eq2}). This expression is lower-bounded by the minimum error
probability of the hypothesis testing problem, denoted by
$\op{Pr}_e(x,x+\tau)$, which does not depend on $\tilde X$, and one
obtains
\begin{align}
\op{Pr}\bk{|\tilde X-X|\ge \frac{\tau}{2}}
&\ge \int_{-\infty}^{\infty} dx [P_X(x)+P_X(x+\tau)] \op{Pr}_e(x,x+\tau).
\end{align}
The left-hand side is a monotonically decreasing function of $\tau$,
so a tighter bound can be obtained if we fill the valleys of the
right-hand side as a function of $\tau$. Denoting this valley-filling
operation as $\mathcal V$:
\begin{align}
\mathcal V f(\tau) &\equiv \max_{\eta \ge 0} f(\tau+\eta),
\end{align}
one gets
\begin{align}
\op{Pr}\bk{|\tilde X-X|\ge \frac{\tau}{2}}
&\ge \mathcal V\int_{-\infty}^{\infty} dx
[P_X(x)+P_X(x+\tau)]\op{Pr}_e(x,x+\tau),
\\
\Sigma &\ge \frac{1}{2}\int_0^\infty d\tau \tau
\mathcal V\int_{-\infty}^{\infty} dx
[P_X(x)+P_X(x+\tau)]\op{Pr}_e(x,x+\tau).
\end{align}
This is a Ziv-Zakai bound. Another version that relates the
mean-square error to an equally-likely-hypothesis-testing problem can
be obtained from Eq.~(\ref{eq1}):
\begin{align}
\op{Pr}\bk{|\tilde X-X|\ge \frac{\tau}{2}}
&\ge \int_{-\infty}^{\infty} dx 2\min[P_X(x),P_X(x+\tau)]
\nonumber\\&\quad\times
\BK{\frac{1}{2}\op{Pr}\bk{\tilde X> x+\frac{\tau}{2}\bigg|X = x}
+\frac{1}{2}\op{Pr}\bk{\tilde X\le x+\frac{\tau}{2}\bigg|X = x+\tau}}.
\end{align}
The expression in the curly brackets is now the error probability of
the same hypothesis testing problem as before, except that the prior
probabilities are
\begin{align}
\op{Pr}(\mathcal H_0)=
\op{Pr}(\mathcal H_1) = \frac{1}{2}.
\end{align}
Another Ziv-Zakai bound follows:
\begin{align}
\Sigma &\ge \frac{1}{2}\int_0^\infty d\tau \tau\mathcal V\int_{-\infty}^{\infty} dx 2\min[P_X(x),P_X(x+\tau)]
\op{Pr}_{e}^{el}(x,x+\tau),
\end{align}
where $\op{Pr}_e^{el}(x,x+\tau)$ now denotes the minimum error
probability with equally likely hypotheses. Note that these bounds
make no assumption about the estimate.
\subsection{\label{qzzb_gauss}Quantum Ziv-Zakai bound for
approximately Gaussian photon-number distributions}
Consider the QZZB for optical phase estimation with
a uniform prior window:
\begin{align}
\Sigma_Z &\equiv \frac{1}{2}\int_0^W d\tau \tau\bk{1-\frac{\tau}{W}}
\Bk{1-\sqrt{1-F(\tau)}},
\\
F(\tau) &\equiv \Big|\sum_n |C_n|^2\exp(in\tau)\Big|^{2},
\label{fourier}
\end{align}
where $|C_n|^2$ is the photon-number distribution. If $|C_n|^2$ can be
approximated as a Gaussian distribution and the sum in
Eq.~(\ref{fourier}) as a continuous Fourier transform,
\begin{align}
|C_n|^2 &\approx \frac{1}{\sqrt{2\pi}\Delta N_j}
\exp\Bk{-\frac{(n-N_j)^2}{2\Delta N_j^2}},
\\
F (\tau) &\approx \Big|\int dn |C_n|^2\exp(in\tau)\Big|^{2}
= \exp\bk{-\Delta N_j^2\tau^2}.
\end{align}
Since $1-\sqrt{1-F} \ge F/2$,
\begin{align}
\Sigma_Z &\gtrsim \frac{1}{4}\int_0^W d\tau \tau\bk{1-\frac{\tau}{W}}
F(\tau)
= \frac{1}{8\Delta N_j^2}
\textrm{ for } W\to\infty,
\end{align}
which is lower than the quantum Cram\'er-Rao bound (QCRB) by a constant
factor of 2. This shows that the two bounds can differ significantly
only when $|C_n|^2$ cannot be well approximated by a Gaussian.
For multiple copies, the fidelity can be written as
\begin{align}
F(\tau) &= \abs{\sum_{n_1,\dots,n_\nu}
|C_{n_1}|^2|C_{n_2}|^2\dots |C_{n_\nu}|^2\exp\bigg(i\sum_j n_j \tau\bigg)}^{2},
\end{align}
which is the squared magnitude of the Fourier transform with respect
to the total photon number $\sum_j n_j$. By virtue of the central
limit theorem, the total-photon-number statistics will become
approximately Gaussian with variance $\nu\Delta N_j^2$ in the limit of
large $\nu$. This means that, regardless of the form of $|C_n|^2$, the
QZZB and the QCRB will become comparable in the limit of large $\nu$.
\subsection{\label{qzzb_coh}Quantum Ziv-Zakai bound for coherent states}
With the inequalities
\begin{align}
\tau\bk{1-\frac{\tau}{W}} &\ge \frac{W}{4}\sin\frac{\pi \tau}{W},
&
0&\le \tau\le W,
\\
1-\sqrt{1-F} &\ge \frac{1}{2}F,
&
0 &\le F \le 1,
\end{align}
the QZZB becomes
\begin{align}
\Sigma_Z &\equiv \frac{1}{2}\int_0^W d\tau \tau\bk{1-\frac{\tau}{W}}
\Bk{1-\sqrt{1-F(\tau)}}
\\
&\ge \frac{W}{16}\int_0^W d\tau \sin\frac{\pi \tau}{W}F(\tau).
\label{lowerZ}
\end{align}
For a coherent state and $W = 2\pi$,
\begin{align}
\Sigma_Z &\ge \frac{\pi}{8}\exp(-2N)
\int_0^{2\pi} d\tau \sin\frac{\tau}{2}\exp\bk{2N\cos \tau}.
\end{align}
Changing the integration variable to $u \equiv \cos(\tau/2)$ and using
the identity $\cos \tau = 2\cos^2(\tau/2)-1$, one obtains
\begin{align}
\Sigma_Z &\ge
\frac{\pi}{2}\exp(-4N)
\int_{-1}^{1} du \exp(4N u^2),
\end{align}
which leads to Eq.~(16) in the text.
\subsection{\label{qzzb_rect}Quantum Ziv-Zakai bound for rectangle states}
Using Eq.~(\ref{lowerZ}) with $W = 2\pi$, one obtains
\begin{align}
\Sigma_Z &\ge \frac{\pi}{8(M+1)^2} I_{M},
\\
I_M &\equiv \int_0^{2\pi} d\tau \frac{\sin^2(M+1)\tau/2}{\sin \tau/2}.
\end{align}
It can be shown using trigonometric identities that
\begin{align}
\sin^2\frac{(M+1)\tau}{2}
&= \sin^2\frac{(M-1)\tau}{2} +
4\sin\frac{\tau}{2}\cos\frac{M\tau}{2}\sin\frac{(M-1)\tau}{2}
+4\sin^2\frac{\tau}{2}\cos^2\frac{M\tau}{2}.
\end{align}
The integral $I_M$ then satisfies the recursive relation
\begin{align}
I_M &= I_{M-2} + \frac{16M}{(2M-1)(2M+1)}
\\
&= I_{M-2} + 4\Bk{\frac{1}{2(M-1)+1}+\frac{1}{2M+1}},
\end{align}
with $I_{-1} = 0$ and $I_{0} = 4$. Hence
\begin{align}
I_M &= 4\sum_{k=0}^M\frac{1}{2k+1}.
\end{align}
For large $M$, the discrete sum can be approximated by an integral:
\begin{align}
I_M &\approx 4\int_0^M dk\frac{1}{2k+1} = 2\ln(2M+1).
\end{align}
\end{widetext}
\end{document}
|
\section{\@startsection {section}{1}{\z@}{-3.5ex plus -1ex minus
-.2ex}{2.3ex plus .2ex}{\large\bf}}
\makeatother
\def\begin{equation}{\begin{equation}}
\def\end{equation}{\end{equation}}
\def\partial{\partial}
\def\displaystyle{\displaystyle}
\begin{document}
\thispagestyle{empty}
\title{
Carleman Estimates and
null controllability of coupled degenerate systems}
\author{E. M. Ait Ben Hassi, \,F. Ammar Khodja, \, A. Hajjaj, \, L. Maniar}
\date{\today}
\subjclass[2000]{35K05, 35K65, 47D06, 93C20}
\begin{abstract} In this paper, we study the null
controllability of weakly degenerate
coupled parabolic systems with two different diffusion coefficients and one control force. To obtain this aim, we develop first new global Carleman estimates for degenerate parabolic equations with weight functions different from the ones of \cite{bouss},
\cite{CanMarVan2008}
and \cite{Martinez1}.
\end{abstract}
\keywords{semigroups, Carleman estimates,
degenerate, parabolic equations, coupled systems, control force, observability inequality,
null controllability.}
\maketitle
\section{Introduction}
This paper is concerned with the null
controllability for the coupled degenerate
parabolic systems
\begin{align}
&u_{t}
-(x^{\alpha_1}u_x)_x + b_{11}(t,x) u+b_{12}(t,x) v=h(t,x) 1_{\omega} ,& (t,x)\in
\times(0,1),\label{eq:1}
\\
&v_{t} -(x^{\alpha_2}v_x)_x +b_{22}(t,x) v+b_{21}(t,x) u=0,& (t,x)\in
\times(0,1),
\\
&u(t, 0) = u(t, 1) = \,\,
v(t, 0) = v(t, 1) = 0, & t\in
(0,T),\\
&u(0, x) =u_0(x),\,\,
v(0, x) = v_0( x) ,& x\in (0,1), \label{eq:2}
\end{align}
where $\omega =(a,b)$ is an open subset of $(0,1)$, $h
\in L^2((0, T)\times[0, 1))$, $(u_0,\, v_0) \in L^2 (0,1)\times L^2 (0,1) $,
$(\alpha_1,\alpha_2)\in(0,1)^2$
and $b_{ij}\in L^\infty((0,T)\times(0,1)), \,\, i,j=1,2$.
Controllability properties of nondegenerate parabolic equations have been widely studied, see
\cite{Cab_Men_Zua}, \cite{DeT_Zua}, \cite{FaPuZu}, \cite{FaRu1}, \cite{FaRu2}, \cite{Cara}, \cite{Cara_Zua}, \cite{Leb_Rob}, \cite{Lo_Zh_Zu}, \cite{Mi_Zu}, \cite{Rus}, \cite{Tat}, using several techniques in particular the Carleman estimates. In \cite{bouss}, \cite{CanMarVan2008}, \cite{Martinez1} new Carleman estimates were developed for degenerate parabolic equations and used to show observability inequalities of the adjoint degenerate problems and then obtain the null controllability. Recently, in \cite{CaToYa} Cannarsa et al. established a local Carleman estimate and deduced unique continuation and boundary approximate controllability for weakly degenerate equations.
The null controllability of coupled parabolic systems was studied for example in
\cite{Khoudja1}, \cite{Khoudja2}, \cite{De0}, \cite{Burgos1}, \cite{Burgos2}, \cite{Gu} in the nondegenerate case.
In \cite{Liu}, Liu et al. considered parabolic cascade systems, $b_{12}=0$, with degeneracy in only
one equation, using the nondegenerate Carleman estimate of Fursikov and Imanuvilov
\cite{Fursikov} and an approximation argument as in \cite{CaMaVa}. In
\cite{Ca-Te}, Cannarsa and De Teresa studied the null controllability of cascade degenerate linear systems with the same diffusion coefficient,
i.e., $\alpha_1=\alpha_2$, and with the particular coupling term $b_{21}=1_O$ for some open set $O\subset (0,1)$.
In \cite{bahm}, we studied the null controllability for degenerate cascade systems with general coupling terms and two different diffusion coefficients. We used a Carleman estimate from \cite{bouss}, and chose carefully appropriate parameters in the weight functions
$\varphi_1(t,x)= \frac{\lambda_1(x^{2-{\alpha_1}}-d_1)}{t^4(T-t)^4}$ and $\varphi_2(t,x)= \frac{\lambda_2(x^{2-{\alpha_2}}-d_2)}{t^4(T-t)^4}$ to obtain the inequality $e^{s\varphi_1}\leq C e^{s\varphi_2}$ to absorb the coupling term.
For general degenerate systems \eqref{eq:1}-\eqref{eq:2}, we need the uniform equivalence $e^{s\varphi_1} \equiv e^{s\varphi_2}$. But this occurs if and only if $\alpha_1=\alpha_2$. To overcome this problem we propose in this paper a common weight function
$\varphi(t,x)= \frac{\lambda(x^{2-{\beta}}-d)}{t^k(T-t)^k}$ for some $\beta$ in terms of $\alpha_1$ and $\alpha_2$.
Then, the first step in this paper is to show new Carleman estimates for the following degenerate parabolic equation
\begin{align}
& y_{t} -(x^{\alpha}y_x)_x = f(t,x), \,\,\,\,\,
& (t,x)\in (0,T)\times(0,1), \label{1eq1} \\
& y(t,0)=y(t,1)=0, \,\,
& t\in(0,T), \\
& y(0,x)=y_0,
& x\in (0,1), \label{1eq2}
\end{align}
with the weight function $\varphi(t,x)= \frac{\lambda(x^{2-{\beta}}-d)}{t^k(T-t)^k}$ with
$d,\lambda$ and $k$ constants to be specified later. To prove our Carleman estimates, we need to show the following fundamental Hardy-Poincar\'e inequality
\begin{align}\label{Hardy1}
\int_0^1x^{\gamma-2}v^2dx\leq C_\gamma \int_0^1x^\gamma v_x^2dx \quad\quad \mbox{where}\,\,\,\,
C_\gamma=\frac{4}{(1-\gamma)^2}
\end{align}
for $\gamma <1$, and $v$ satisfying $v(0)=0$ and $\int_0^1x^\gamma v_x^2dx <+\infty$. This result was proved in \cite{bouss}, \cite{CanMarVan2008} and \cite{Martinez1} for $0<\gamma<2, \gamma\neq 1$. But, for our Carleman estimates we need this inequality for negative $\gamma$, see Lemma \ref{Hardy2}.
This will allow us to deduce Carleman estimates for the adjoint coupled degenerate system
\begin{align}
& U_{t} -(x^{\alpha_1}U_x)_x + b_{11}(t,x) U+b_{21}(t,x) V=0, & (t,x)\in (0,T)\times(0,1), \label{adjfirst}\\
& V_{t}-(x^{\alpha_2}V_x)_x +b_{22}(t,x) V+ b_{12}(t,x) U=0, & (t,x)\in (0,T)\times(0,1),\label{21}\\
& U(t,1)=U(t,0)=\, V(t,1)=V(t,0)=0 ,& t\in (0,T),\label{23}\\
& U(0,x)=U_0(x), V(0,x)=V_0(x), & x\in (0,1)\label{adjend},
\end{align}
and then its observability inequality. Using a standard argument, we obtain the null controllability of \eqref{eq:1}-\eqref{eq:2}. By a linearization argument and fixed point, see for example \cite{bahm}, \cite{bouss}, \cite{Can-Gen-2006}, \cite{Zuazua99} one can show easily the null controllability of semilinear degenerate coupled systems.
This paper is organized as follows. Section 2 is devoted to the well-posedness of the coupled degenerate systems. In section 3, we establish our new Carleman estimates for degenerate parabolic equations and deduce similar estimates for the coupled degenerate systems.
In section 4, we deduce observability inequality and null controllability results. In appendix, we give summarized proofs of Caccioppoli and Hardy-Poincar\'e inequalities.
\section{Well-posedness}
In order to study the well-posedness of the system
\eqref{eq:1}-\eqref{eq:2}, we introduce the weighted spaces
$$
H_{\alpha_i}^1(0, 1)
:=\Big\{ u \in L^2(0, 1) :\, u \,\mbox{ is abs. continuous in}\,
[0, 1],\,\, x^{\alpha_i/2} u_x \in L^2(0, 1)\, \mbox{and}\, u(0)=u(1) = 0\Big \}
$$
with the norm
$\|u\|^2_{ H^1_{\alpha_i}(0, 1)} := \|u\|^2_{L^2(0,1)} +
\|x^{\alpha_i/2}u_x\|^2_{ L^2(0,1)}$ and
$$H^2_{\alpha_i} (0, 1):=\Big\{ u \in H^1_{\alpha_i}(0, 1) \,: \, x^{\alpha_i}u_x
\in H^1(0, 1)\Big\}$$
with the norm $$\|u\|^2_{H^2_{\alpha_i}(0, 1)} := \|u\|^2 _{H^1_{\alpha_i}(0, 1)}
+ \|(x^{\alpha_i}u_x)_x\|^2_{ L^2(0,1)}.$$
We define the operator $(A_i,D(A_i))$ by
$$
A_iu := (x^{\alpha_i}u_x)_x ,\, \, \, u \in D(A_i) = H^2_{\alpha_i}(0, 1), \; i=1,2.
$$
We recall the following properties of $(A_i,D(A_i))$.
\begin{proposition} (\cite {cmp}, \cite{CaMaVa}). For $i=1,2$, the operator $A_i : D(A_i) \longrightarrow L^2(0, 1)$ is closed,
self-adjoint, negative and with dense domain.
\end{proposition}
In the Hilbert space $\mathbb{H}:= L^2(0,1) \times L^2(0, 1) $,
the system \eqref{eq:1}-\eqref{eq:2} can be
transformed in the following Cauchy problem
$$
(CP)\quad
\begin{cases} X'(t) =\mathcal{A}X(t) +B(t)X(t)+G(t),\\
X(0)=\left(\begin{smallmatrix}
u_0\\
v_0
\end{smallmatrix}\right),
\end{cases}
$$ where $X(t)=\left(\begin{smallmatrix}
u(t)\\
v(t)
\end{smallmatrix}
\right),$ $\mathcal{A}=\left(\begin{matrix}
A_1&0\\
0&A_2
\end{matrix}
\right) $, $D(\mathcal{A})=D(A_1)\times D(A_2), \,\, G(t)=\left(\begin{smallmatrix}
h(t,x) 1_\omega\\
0 \end{smallmatrix}
\right)$,
and
$$
B(t)=\left(\begin{matrix}
M_{b_{11}(t)}&M_{b_{12}(t)}\\
M_{b_{21}(t)}&M_{b_{22}(t)}
\end{matrix}
\right), \,\, \mbox{where}\,\, M_{b_{ij}(t)}u=b_{ij}(t)u.
$$
As the operator $\mathcal{A}$
is diagonal and since $B(t)$ is a bounded perturbation, the following wellposedness and regularity
results hold.
\begin{proposition}\label{estimsemigroup}
(i) \, The operator $\mathcal{A}$ generates a contraction strongly
continuous semigroup $(T(t))_{t\geq0}.$
\\
(ii)\, For all $h\in L^2((0,T)\times (0,1))$ and $(u_0,v_0)\in L^2(0,1)\times L^2(0,1)$ there exists a unique mild solution $(u,v)\in X_T:=C\left( [0,T],L^2(0,1)\times L^2(0,1) \right)
\cap L^2\left(0,T;H^1_{\alpha_1}\times H^1_{\alpha_2}\right)$ of \eqref{eq:1}-\eqref{eq:2} satisfying
\begin{align}\label{estiminitialdata}
\sup_{[0,T]}\|(u,v)(t)\|^2_{L^2\times L^2} + \int_0^T \| (x^{\frac{\alpha_1}{2}}u_x ,x^{\frac{\alpha_2}{2}}v_x) \|^2_{L^2} \, dt
\nonumber \\
\leq C_T\left(\|(u_0,v_0)\|^2_{L^2\times L^2}+
\|h\|^2_{L^2((0,T)\times (0,1))}\right)
\end{align}
for a constant $C_T>0$.
Morover, if $(u_0,v_0)\in H^1_{\alpha_1}\times H^1_{\alpha_2}$ then, $(u,v)\in Y_T:=C\left([0,T],H^1_{\alpha_1}\times H^1_{\alpha_2}\right)
\cap H^1\left(0,T;L^2(0,1)\times L^2(0,1)\right)\cap L^2\left(0,T;H^2_{\alpha_1}\times H^2_{\alpha_2}\right )$ and
$$
\sup_{[0,T]}\|(u,v)(t)\|^2_{H^1_{\alpha_1}\times H^1_{\alpha_2}} + \int_0^T \left( \| (u_t,v_t) \|^2_{L^2} + \| ((x^{\alpha_1}u_x)_x,(x^{\alpha_2}v_x)_x) \|^2_{L^2} \right) \, dt $$
$$
\leq C_T\left(\|(u_0,v_0)\|^2_{H^1_{\alpha_1}\times H^1_{\alpha_2}}+
\|h\|^2_{L^2((0,T)\times (0,1))}\right)
$$
for a constant $C_T>0$.
\end{proposition}
\section{Carleman estimates}
In this section we prove new Carleman estimates for the adjoint system \eqref{adjfirst}-\eqref{adjend}.
For this, let
$\omega':=(a',b')\Subset \omega$
and let
us introduce the weight functions :
$\varphi(t,x):=\Theta(t)\psi(x)$;
\,\, $\Theta(t):=\dfrac{1}{t^k(T-t)^k};$ \,\, $\psi(x):=\lambda\left(x^{2-\beta}-d\right)$; \quad
$\Phi(t,x)=\Psi(x)\Theta(t)$; \quad $\Psi (x):=\left(e^{\rho \sigma(x)}-e^{2\rho||\sigma||_\infty}\right)$; \quad
$\phi(t,x)=e^{\rho\sigma(x)}\Theta(t)$;
where $\sigma$ is a function in $C^2([a',1])$
satisfying $\sigma(x)>0$ in $(a',1)$, $\sigma(a')=\sigma(1)=0$ and $\sigma_x(x)\neq 0$ in
$[a',1]\backslash\omega_0$ for some open $\omega_0\Subset (a',1)$ and the parameters $d$, $\rho$, $\lambda$ and $k$ are chosen such that $d\geq 5$; $\rho > \frac{4ln 2}{||\sigma||_\infty}$, $\frac{e^{2\rho||\sigma||_\infty}}{d-1} < \lambda < \frac{4}{3d}(e^{2\rho||\sigma||_\infty}-e^{\rho||\sigma||_\infty})$ and $k\geq 4$.
\begin{remark}\label{rk1}
\begin{itemize}
\item
These weight functions are independent of the diffusion coefficient. This play a crucial role to study coupled system of non cascade form.
\item
The existence of the function $\sigma$ was proved for example in \cite{Fursikov} using Morse functions. But in 1-dimension one can show this easily using cut-off functions.
\item
If $d\geq 5$ and $\rho > \frac{4ln 2}{||\sigma||_\infty}$ then the interval $\left]\frac{e^{2\rho||\sigma||_\infty}}{d-1}\, , \frac{4}{3d}(e^{2\rho||\sigma||_\infty}-e^{\rho||\sigma||_\infty})\right[$ is not empty. We can then choose $\lambda$ in this interval.
\item
For this choice of the parameters $d,\, \rho$ and $\lambda $ the weight functions $\varphi$ and $\Phi$
satisfy the following inequalities which are needed in the sequel
\begin{align}\label{comparaison phi}
\frac43 \Phi < \varphi < \Phi \,\, \mbox{on } (0,T)\times(0,1).
\end{align}
\item
For nondegenerate problems one needs the following estimates see e.g. \cite{Fursikov}
\begin{align}\label{majorTheta1}
\lim_{t\rightarrow O^+} \Theta(t)=\lim_{t\rightarrow T^-} \Theta(t)= +\infty,\quad
\Theta(t)\geq c_1, \quad |\dot{\Theta}|\leq c_2 \Theta^2, \quad |\ddot{\Theta}|\leq c_3 \Theta^3.
\end{align}
and this is satisfied for all $k\geq 1$ with $c_1=(2/T)^{2k}$, $c_2=kT(T/2)^{2(k-1)}$, $c_3=k(k+1)T^2(T/2)^{4(k-1)}$.
\item
For the degenerate case one needs in addition the estimate
\begin{align}\label{majorTheta2}
|\ddot{\Theta}|\leq c_4 \Theta^{2}.
\end{align}
which is satisfied for all $k\geq 2$ with \textcolor[rgb]{0.98,0.00,0.00}{$c_4=k(k+1)T^2(T/2)^{k-4}$}.
\end{itemize}
\end{remark}
We begin by proving first a new Carleman estimate for the problem \eqref{1eq1}-\eqref{1eq2} with one equation.
\begin{theorem}\label{carl_1eq}
Let $T>0$ and suppose that $y_0\in H^1_{\alpha}$. Then, for all $\beta\in [\alpha,1)$
there exist two positive constants $C$ and $s_0$ such that every solution $y$ of
\eqref{1eq1}-\eqref{1eq2} satisfies for all $s\geq s_0$
\begin{align}
& \int_0^T\int_0^1\left( s\Theta(t) x^{2\alpha-\beta}y_x^2 + s^3 \Theta^3(t)
x^{2+2\alpha-3\beta}y^2 \right)e^{2s\varphi (t,x)}\,dx dt \notag \\
& \vspace*{2cm}\leq C\left( \int_0^T\int_0^1 f^2(t,x) e^{2s\varphi (t,x)}\,dxdt+ \int_0^T s\Theta(t)y_x^2(t,1)e^{2s\varphi (t,1)}dt\right).\quad \quad \label{carl1equ}
\end{align}
\end{theorem}
\begin{proof}
For $s>0$, let us introduce the function $z:=e^{s\varphi}y$.
We have $$L_sz:=z_t+(x^\alpha z_x)_x -2sx^\alpha \varphi_xz_x -s\varphi_tz+s^2x^\alpha \varphi_x^2 z - s(x^\alpha \varphi_x)_x z =fe^{s\varphi}.$$
Let
\begin{align*}
& L_s^+z := (x^\alpha z_x)_x-s\varphi_tz+s^2x^\alpha \varphi_x^2 z, \\
& L_s^-z := z_t-2sx^\alpha \varphi_xz_x - s(x^\alpha \varphi_x)_x z, \\
& \hspace*{.35cm} f_s := fe^{s\varphi}.
\end{align*}
We have
$||f_s||_{L^2}^2=||L_s^+z+L_s^-z ||_{L^2}^2=||L_s^+z ||_{L^2}^2+||L_s^-z ||_{L^2}^2+2\langle L_s^+z,L_s^-z\rangle \geq 2\langle L_s^+z,L_s^-z\rangle$. One has $z(0,x)=z(T,x)=z_x(0,x)=z_x(T,x)=0$.
So integrating by parts one obtains
$$\langle L_s^+z,L_s^-z\rangle= -2s^2\int_0^T\int_0^1 x^\alpha \varphi_x \varphi_{tx}z^2 \,dx\,dt
+ s\int_0^T\int_0^1 x^\alpha(x^\alpha \varphi_x)_{xx}zz_x \,dx\,dt $$
$$\frac{s}{2}\int_0^T\int_0^1 \varphi_{tt}z^2 \,dx\,dt
+s\int_0^T\int_0^1 x^\alpha \left[2x^\alpha\varphi_{xx}+\alpha x^{\alpha-1}\varphi_x \right]z_x^2 \,dx\,dt$$
$$+s^3 \int_0^T\int_0^1 x^\alpha \left[2x^\alpha\varphi_{xx}+\alpha x^{\alpha-1}\varphi_x \right]\varphi_x^2z^2
$$
$$ + \int_0^T \left[
x^\alpha z_xz_t
+s^2\Theta \dot{\Theta}\psi\psi_x x^\alpha z^2
- s^3\Theta^3 x^{2\alpha}\psi_x^3 z^2
\right]_{x=0}^{x=1} dt
$$
$$ - \int_0^T \left[ \lambda(2-\beta)s\Theta x^{1-\beta}(x^\alpha z_x)^2
+ \lambda(2-\beta)(1+\alpha-\beta) s\Theta x^{2\alpha-\beta}zz_x \right]_{x=0}^{x=1} dt.
$$
It is easy to check that if $y\in H^2_{\alpha}(0,1)$ then we have also $z\in H^2_{\alpha}(0,1).$ So $x^\alpha z \in H^1(0,1)\subset L^\infty (0,1)$ by the Sobolev imbedding theorem. Then, using the facts that $z(t,0)=z(t,1)=z_t(t,0)=z_t(t,1)=0$ and $x^\alpha z_x$, $x^\alpha$, $\psi, \, \psi_x$ are bounded, we deduce that the first integral with boundary terms vanishes
and $x^{1-\beta}(x^\alpha z_x)^2|_{x=0}=0$.
On the other hand we have $\left[ x^{2\alpha-\beta}zz_x \right]_{x=0}^{x=1}=0$,
in fact it is clear that $x^{2\alpha-\beta}zz_x|_{x=1}=(x^\alpha z_x)z|_{x=1} = 0$
and since $x^\alpha z_x \in L^\infty (0,1)$ and $z(t,0)=0$
then for each $t\in(0,T)$ we have
\begin{align}\label{mojor z,zx}
|z_x(t,x)|\leq c x^{-\alpha} \,\, \mbox{and} \,\,\, |z(t,x)| =|\int_0^x z_x(t,y)dy|\leq c x^{1-\alpha}.
\end{align}
Therefore $|x^{2\alpha-\beta}zz_x(t,x)|\leq c x^{1-\beta}$. Consequently, since $\beta<1$ we deduce $x^{2\alpha-\beta}zz_x|_{x=0}=0$.
\\
We have then
$$\underset{J_1}{\underbrace{\lambda^3(2-\beta)^3(2-2\beta+\alpha)\int_0^T\int_0^1s^3\Theta^3x^{2+2\alpha-3\beta}z^2\, dxdt}}
+ \underset{J_2}{\underbrace{\lambda(2-\beta)(2-2\beta+\alpha)\int_0^T\int_0^1s\Theta x^{2\alpha-\beta}z_x^2 \, dxdt}}$$
$$\leq \frac12 \int_0^T\int_0^1 f^2e^{2s\varphi}\, dxdt
-\underset{J_3}{\underbrace{2\lambda^2(2-\beta)^2\int_0^T\int_0^1s^2\Theta\dot{\Theta} x^{2+\alpha-2\beta}z^2 \,dxdt}}$$
$$+ \underset{J_4}{\underbrace{\lambda(2-\beta)(1+\alpha-\beta)(\beta -\alpha)\int_0^T\int_0^1s\Theta x^{2\alpha-\beta-1}zz_x \, dxdt}}\,
+ \underset{J_5}{\underbrace{\frac{\lambda}{2}\int_0^T\int_0^1s\ddot{\Theta}(d-x^{2-\beta})z^2\,dxdt}} $$
$$+ \lambda(2-\beta)\int_0^Ts\Theta z_x^2 (t,1)dt.$$
Now we will show that $J_3, \, J_4$ and $J_5$ can be absorbed by $J_1$ and $J_2$. For this, let $\varepsilon>0$ fixed to be specified later. First, Since $\beta\geq \alpha$ and $|\Theta\dot{\Theta}|\leq C\Theta^3$ then
$$|J_3|\leq C \int_0^T\int_0^1s^2\Theta^3 x^{2+2\alpha-3\beta}z^2\, dxdt\leq \varepsilon J_1$$
for $s$ large enough.
In the other hand for $J_4$ we have
\begin{align}
|J_4|
& \leq \lambda(2-\beta)(1+\alpha-\beta)(\beta-\alpha)\int_0^T\int_0^1\left[\sqrt{s\Theta} x^{\alpha-\frac{\beta}{2} -1}|z|\right]\left[\sqrt{s\Theta}x^{\alpha-\frac{\beta}{2}}|z_x|\right]\, dxdt \nonumber \\
& \leq \lambda(2-\beta)(1+\alpha-\beta)(\beta-\alpha)\left(\varepsilon\int_0^T\int_0^1 s\Theta x^{2\alpha-\beta -2}z^2dxdt + \frac{1}{4\varepsilon}\int_0^T\int_0^1 s\Theta x^{2\alpha-\beta }z_x^2 dxdt\right)
\label{2341}
\end{align}
Now we will use the Hardy-Poincar\'e inequality \eqref{Hardy}. We have $2\alpha-\beta<1$ and we will show that $\int_0^1x^{2\alpha-\beta} z_x^2dx <+\infty$.
Using \eqref{mojor z,zx} and the fact that $\beta<1$ we obtain,
$$|x^{2\alpha-\beta} z_x^2|
\, \leq \, C x^{-\beta} \in L^1(0,1).$$
We have then
$$\int_0^1x^{2\alpha-\beta-2} z^2dx \leq C_{2\alpha-\beta} \int_0^1x^{2\alpha-\beta} z_x^2dx$$ where $C_{2\alpha-\beta}= \frac{4}{(1-2\alpha+\beta)^2}$.
Then, we get from \eqref{2341}
$$|J_4| \leq \lambda(2-\beta)(1+\alpha-\beta)(\beta-\alpha)\left( \varepsilon C_{2\alpha-\beta}+\frac{1}{4\varepsilon}\right)\int_0^T\int_0^1 s\Theta x^{2\alpha-\beta }z_x^2 dxdt$$
The quantity $\varepsilon C_{2\alpha-\beta}+\frac{1}{4\varepsilon}$ is minimal for $\varepsilon = \frac{1}{2.\sqrt{C_{2\alpha-\beta}}}$. For this choice we have
\begin{align*}
|J_4| \,\leq \, \lambda(2-\beta)
(1+\alpha-\beta)(\beta-\alpha)\frac{2}{1-2\alpha+\beta}
\int_0^T\int_0^1 s\Theta x^{2\alpha-\beta }z_x^2 dxdt
\end{align*}
and for all $\beta\in [\alpha,1)$ we have
\begin{align*}
\frac{2(1+\alpha-\beta)(\beta-\alpha)}{1-2\alpha+\beta} -(2-2\beta+\alpha)=\frac{(\beta-1)(2-\alpha)}{1-2\alpha+\beta} < 0
\end{align*}
The term $J_4$ can then be absorbed by $J_2$.\\
For the last term $J_5$, since
$|\ddot{\Theta}|\leq c_4\Theta^{2}$ and $\beta \geq \alpha$, we have by applying the Hardy-Poincar\'e inequality
\begin{align}
|J_5| & \leq \lambda d c_4 \int_0^T\int_0^1s\Theta^2 z^2dxdt
\nonumber
\\
& = \lambda d c_4\int_0^T\int_0^1\left[\sqrt{s\Theta} x^{\alpha-\frac{\beta}{2} -1}z\right]
\left[\sqrt{s}\Theta^{\frac32} x^{1-\alpha+\frac{\beta}{2}}z\right]\, dxdt \nonumber\\
& \leq \lambda d c_4\int_0^T\int_0^1 \left(\varepsilon s\Theta x^{2\alpha-\beta -2}z^2 +
\frac{1}{4\varepsilon} s\Theta^3 x^{2-2\alpha+\beta }z^2\right) dxdt \nonumber\\
& \leq
\varepsilon \lambda d c_4C_{2\alpha-\beta} \int_0^T\int_0^1 s\Theta x^{2\alpha-\beta}z_x^2\, dxdt
+
C_\varepsilon \int_0^T\int_0^1 s\Theta^3 x^{2+2\alpha-3\beta }z^2\, dxdt
\nonumber
\end{align}
Therefore by choosing $\varepsilon$ small enough, we obtain
$$\int_0^T\int_0^1s^3\Theta^3x^{2+2\alpha-3\beta}z^2\, dxdt
+ \int_0^T\int_0^1s\Theta x^{2\alpha-\beta}z_x^2 \, dxdt$$
$$\leq C \left( \int_0^T\int_0^1 f^2e^{2s\varphi}\, dxdt
+ \int_0^Ts\Theta z_x^2 (t,1)dt\right).$$
for $s$ large enough. So replacing $z$ by $e^{s\varphi}y$ we deduce immediately the conclusion of the theorem.
\end{proof}
\begin{theorem}\label{carl22_1eq}
Let $T>0$ and suppose that $y_0\in H^1_\alpha$.
Then, for all $\beta\in [\alpha,1)$ there exist two positive
constants $C$ and $s_0$ such that every solution $y$ of
\eqref{1eq1}-\eqref{1eq2} satisfies for all $s\geq s_0$
\begin{align}
& \int_0^T\int_0^1\left( s\Theta(t) x^{2\alpha-\beta}y_x^2 + s^3 \Theta^3(t)
x^{2+2\alpha-3\beta}y^2 \right)e^{2s\varphi (t,x)}\,dx dt \notag \\
& \vspace*{2cm}\leq C\left( \int_0^T\int_0^1 f^2(t,x) e^{2s\Phi (t,x)}\,dxdt+
\int_0^T \int_{\omega'}
s^3\phi^3 y^2 e^{2s\Phi (t,x)} dxdt\right)
\label{estimglobal}
\end{align}
\end{theorem}
\begin{proof}
Let us consider an arbitrary open subset $\omega'':=(a'',b'') \Subset\omega'$ and a cut-off function $\xi\in \mathcal{C}^\infty(0,1)$
such that
$$
\begin{cases} 0\leq \xi(x)\leq 1, &x\in(0,1),\\
\xi(x)=1 ,&0\leq x\leq a'',\\
\xi(x)=0 , & b''\leq x\leq 1.
\end{cases}
$$
Let $z=\xi y$ where $y$ is the solution of
\eqref{1eq1}-\eqref{1eq2}. Then $z$ satisfies
the following system
\begin{align}
& z_t-(x^\alpha z_x)_x= \xi f -\xi_x x^\alpha y_x-(x^\alpha\xi_xy)_x, & (t,x)\in(0,T)\times(0,1), \label{firstw}\\
& z(t,1)=z(t,0)=0, & t\in (0,T), \label{tirtw}
\end{align}
Therefore, applying the Carleman estimate \eqref{carl1equ} to the equation \eqref{firstw}
we obtain
\begin{align}\label{es3288}
& \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha-\beta}
z_x^2(t,x)+s^3 \Theta^3(t)x^{2+2\alpha-3\beta}
z^2(t,x)]e^{2s\varphi}dx dt
\notag \\
&\leq C \int_0^T\int_0^1 [\xi^2f^2+\left(\xi_x x^\alpha
y_x+(x^\alpha \xi_xy)_x\right)^2]e^{2s\varphi}dx dt.
\end{align}
So using the definition of $\xi$ and the Cacciopoli's inequality, see Lemma \ref{cacci}, we obtain
\begin{align}\label{es3377}
\int_0^T\int_0^1\left(\xi_x x^\alpha y_x+ (x^\alpha\xi_xy)_x\right)^2e^{2s\varphi}dxdt
&\leq C \int_0^T\int_{\omega''}[y^2+y_x^2]e^{2s\varphi}dx dt\nonumber \\
& \leq C\int_0^T\int_{\omega'} y^2e^{2s\varphi}dx dt.
\end{align}
and
\begin{align}
\int_0^T\int_0^1s\Theta x^{2\alpha-\beta}\xi^2 y_x^2 e^{2s\varphi}dxdt\leq 2 \int_0^T\int_0^1 s\Theta x^{2\alpha-\beta}z_x^2e^{2s\varphi}dxdt + 2\int_0^T\int_{\omega'}s\Theta y^2e^{2s\varphi}dxdt
\end{align}
Thus from \eqref{es3288}-\eqref{es3377} and the definition of $\xi$ we deduce the following estimate
\begin{align}\label{estimleft}
& \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha-\beta}
\xi^2y_x^2(t,x)+s^3 \Theta^3(t)x^{2+2\alpha-3\beta}
\xi^2y^2(t,x)]e^{2s\varphi}dx dt
\notag \\
&\leq C \left( \int_0^T\int_0^1 \xi^2 f^2e^{2s\varphi}dxdt + \int_0^T\int_{\omega'} s\Theta y^2e^{2s\varphi}dxdt\right).
\end{align}
On $(a',1)$ the equation \eqref{1eq1} is uniformly parabolic
hence, one can use the following Carleman estimate which is a consequence of
(\cite{Fursikov}, Lemma 1.2) established by Fursikov and Imanuvilov.
\begin{proposition}
Consider the nondegenerate linear problem
$$\begin{cases}
v_t-(x^\alpha v_x)_x=f \in L^2((0,T)\times(a',1)), \\
v(t,a')=v(t,1)=0,\,\,\, t\in (0,T), & \label{alban-Can2}
\end{cases}$$
Then, there exists a constant $\rho_0 >0$ such that for all $\rho\geq\rho_0$ there exists $s_0(\rho)>0$ such that for each $s\geq s_0(\rho)$ the solution $v$ of the last problem
satisfy the following estimate:
\begin{align}\label{classicFursikov}
& \int_0^T\int_{a'}^1 (s\phi v_x^2+
s^3 \phi^3 v^2)e^{2s\Phi}dx dt \notag \\
& \leq C \left(\int_0^T\int_{a'}^1 f^2e^{2s\Phi}dx dt + \int_0^T\int_{\omega'} s^3 \phi^3 v^2e^{2s\Phi}dx dt\right)
\end{align}
where the functions
$\Phi$ and $\phi$ are defined in Theorem \ref{carl22_1eq}.
\end{proposition}
\begin{remark}
The last estimate was showed in \cite{Fursikov} for $\Theta(t)=\frac{1}{t(T-t)}$
but by careful examination of the proof one can see easily that it remains valid for all $\Theta\in C^2(0,T)$ satisfying
\eqref{majorTheta1}, see Remark \ref{rk1}.
\end{remark}
To achieve the proof of the Theorem \ref{carl01}, let $Z:=\zeta y,$ where the function
$\zeta$ is defined as $\zeta=1-\xi.$ Then $Z$ is a solution of the following problem
\begin{align*}
& Z_t-(x^\alpha Z_x)_x=\zeta f-\zeta_x x^\alpha
y_x-(x^\alpha\zeta_x y)_x, & (t,x)\in(0,T)\times(a',1),
\\
& Z(t,1)=Z(t,a'
)=0 ,& t\in (0,T),
\end{align*}
Applying the classical Carleman estimate \eqref{classicFursikov},
it follows that for $s$ large enough
\begin{align*}
& \int_0^T\int_{0}^1 (s\phi Z_x^2 +
s^3 \phi^3 Z^2)e^{2s\Phi}dx dt \\
& \leq C\left(\int_0^T\int_{0}^1 \left[\zeta f+\zeta_x x^\alpha
y_x+(x^\alpha \zeta_x y)_x\right]^2e^{2s\Phi}dx dt
+ \int_0^T\int_{\omega'} s^3 \phi^3 Z^2e^{2s\Phi}dx dt\right) \\
& \leq C\left(\int_0^T\int_{0}^1 \zeta^2 f^2e^{2s\Phi}dx dt
+ \int_0^T\int_{\omega''}[y^2+y_x^2] e^{2s\Phi}dxdt\right.
\left.+ \int_0^T\int_{\omega'} s^3 \phi^3 Z^2e^{2s\Phi}dx dt \right)
\end{align*}
Therefore, using the Caccioppoli inequality and the definitions of $Z$ and $\zeta$ we deduce
\begin{align}
& \int_0^T\int_{0}^1 (s\phi \zeta^2y_x^2 +
s^3 \phi^3 \zeta^2y^2)e^{2s\Phi}dx dt \nonumber \\
& \leq C\left(\int_0^T\int_{0}^1 \zeta^2 f^2e^{2s\Phi}dx dt
+ \int_0^T\int_{\omega'} s^3 \phi^3 y^2e^{2s\Phi}dx dt \right)\label{estimright}
\end{align}
Thanks to \eqref{comparaison phi} there exists a constant $c>0$ such that for all $(t,x)\in[0,T]\times(a',1)$ one has
\begin{align}\label{compare varphi and Phi}
\Theta x^{2\alpha-\beta}e^{2s\varphi(t,x)}\leq c
\phi e^{2s\Phi(t,x)} \,\mbox{and} \,\,
\Theta^3 x^{2+2\alpha-3\beta}e^{2s\varphi(t,x)}\leq c \phi^3e^{2s\Phi(t,x)}
\end{align}
Then, using \eqref{estimleft}, \eqref{estimright}, \eqref{comparaison phi}, \eqref{majorTheta1} and the fact that $1/2 \leq \xi^2+\zeta^2\leq 1$
we obtain the global estimate
\begin{align}
& \int_0^T\int_0^1\left( s\Theta(t) x^{2\alpha-\beta}y_x^2 + s^3 \Theta^3(t)
x^{2+2\alpha-3\beta}y^2 \right)e^{2s\varphi (t,x)}\,dx dt \notag \\
& \vspace*{2cm}\leq C\left( \int_0^T\int_0^1 f^2(t,x) e^{2s\Phi (t,x)}\,dxdt+
\int_0^T \int_{\omega'} s^3\phi^3 y^2 e^{2s\Phi (t,x)} dxdt\right)
\label{estimglobal1}
\end{align}
This ends the proof of Theorem \ref{carl22_1eq}.
\end{proof}
The estimate in Theorem \ref{carl22_1eq} was obtained for regular initial data.
By density we deduce the following result for the general case: $y_0 \in L^2(0,1)$.
%
%
%
\begin{corollary}
Let $T>0$ be given. Let $\beta\in [\alpha,1)$ and $\mu\geq max(0,2+2\alpha-3\beta)$. Then there exist two positive
constants $C$ and $s_0$ such that every solution $y$ of
\eqref{1eq1}-\eqref{1eq2} satisfies for all $s\geq s_0$
\begin{align}
& \int_0^T\int_0^1\left( s\Theta x^{\alpha}y_x^2 + s^3 \Theta^3
x^{\mu}y^2 \right)e^{2s\varphi (t,x)}\,dx dt \notag \\
& \vspace*{2cm}\leq C\left( \int_0^T\int_0^1 f^2(t,x) e^{2s\Phi (t,x)}\,dxdt+
\int_0^T \int_{\omega'}
s^3\phi^3
y^2
e^{2s\Phi (t,x)}
dxdt\right)
\label{estimgloballll}
\end{align}
\end{corollary}
\begin{proof}
Let $y_0\in L^2(0,1).$ By the density of $H^1_\alpha(0,1)$ in $L^2(0,1)$, there exist a set $(y_0^n)_n$ in $H^1_\alpha(0,1)$ which converges to $y_0$.
Let $y^n$ the unique solution in the space
$Z_T:=C\left( [0,T],L^2(0,1) \right)
\cap L^2\left(0,T;H^1_{\alpha}\right)$
of the problem \eqref{1eq1}-\eqref{1eq2} associated to the initial data $y_0^n$. As in \eqref{estiminitialdata}
one has for a constant $C_T>0$
$$\|(y^m-y^n)(t)\|_{Z_T}:= \sup_{[0,T]}\|(y^m-y^n)(t)\|^2_{L^2} + \int_0^T \| x^{\frac{\alpha}{2}}(y^m-y^n)_x \|^2_{L^2} \, dt
\leq C_T \|y^m_0-y^n_0\|^2_{L^2}.$$
Therefore the set $(y^n)_n$ has a limit $y$ in the Banach space $Z_T$. Using classical argument in semigroup theory it is easy to show that $y$ is the solution of the problem \eqref{1eq1}-\eqref{1eq2} associated to the initial data $y_0$.
On the other hand since
$x^{\alpha}\leq x^{2\alpha-\beta}$ and $x^{\mu}\leq x^{2+2\alpha-3\beta}$ on (0,1) then we deduce from Theorem \ref{carl22_1eq} the estimate
\begin{align*}
& \int_0^T\int_0^1\left( s\Theta x^{\alpha}|y^n_x|^2 + s^3 \Theta^3
x^{\mu}|y^n|^2 \right)e^{2s\varphi (t,x)}\,dx dt \notag \\
& \vspace*{2cm}\leq C\left( \int_0^T\int_0^1 f^2(t,x) e^{2s\Phi (t,x)}\,dxdt+
\int_0^T \int_{\omega'} s^3\phi^3|y^n|^2 e^{2s\Phi (t,x)}dxdt\right)
\end{align*}
And since $s\Theta e^{2s\varphi}$, $s^3\Theta^3 e^{2s\varphi}x^{\mu}$ and $s^3\phi^3 e^{2s\Phi}$ are bounded then one can pass to the limit and get the desired estimate.
\end{proof}
For the coupled system \eqref{adjfirst}-\eqref{adjend} we prove first an intermediate important result which could be used to show the null controllability for a coupled system with two control forces
\begin{theorem}\label{carl01}
Let $T>0$ and $(\alpha_1,\alpha_2)\in (0,1)\times(0,1)$ be given and suppose that $y_0\in H_\alpha^1$. Then for all
$\beta \in [max(\alpha_1,\alpha_2),1[$
there exist two positive
constants $C$ and $s_0$ such that every solution $(U,V)$ of
\eqref{adjfirst}-\eqref{adjend} satisfies
\begin{align}\label{Carl2degenr}
& \int_0^T\int_0^1 s\Theta(t) \left[ x^{2\alpha_1-\beta}U_x^2(t,x)+x^{2\alpha_2-\beta}V_x^2(t,x)\right]e^{2s\varphi (t,x)}\,dx dt \notag \\
& +\int_0^T\int_0^1 s^3 \Theta^3(t)\left[
x^{2+2\alpha_1-3\beta}U^2(t,x)+x^{2+2\alpha_2-3\beta}V^2(t,x)\right] e^{2s\varphi (t,x)}\,dx dt \notag \\
& \vspace*{2cm}\leq C\int_0^T\int_{\omega'} s^3
\Theta^3 \left[U^2(t,x)+V^2(t,x)\right]e^{2s\Phi(t,x)}
\,dxdt \quad \quad \mbox{for all}\,\,s\geq s_0.
\end{align}
\end{theorem}
\begin{proof}
Since $U$ is solution of the problem
\begin{align*}
& U_{t} -(x^{\alpha_1}U_x)_x =- b_{11}(t,x) U-b_{21}(t,x) V, & (t,x)\in (0,T)\times(0,1), \\
& U(t,1)=U(t,0)=0,\, & t\in (0,T),\\
& U(0,x)=U_0(x), & x\in (0,1),
\end{align*}
then applying the estimate \eqref{estimleft} to this system we obtain
\begin{align}\label{estimU}
& \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_1-\beta}
\xi^2U_x^2(t,x)+s^3 \Theta^3(t)x^{2+2\alpha_1-3\beta}
\xi^2U^2(t,x)]e^{2s\varphi}dx dt
\notag \\
&\leq \overline{C} \int_0^T\int_0^1 \xi^2 (b_{11}^2U^2+b_{21}^2V^2)e^{2s\varphi}dxdt + C\int_0^T\int_{\omega'}s\Theta U^2e^{2s\varphi}dxdt.
\end{align}
Using the Hardy-Poincar\'e inequality \eqref{Hardy} one has for $s$ large enough
\begin{align*}
\int_0^T\int_0^1 b_{11}^2 \xi^2U^2e^{2s\varphi}dxdt &\leq C
\int_0^T\int_0^1\left[ x^{\alpha_1 -\frac{\beta}{2} -1}\xi Ue^{s\varphi}\right]
\left[ x^{1-\alpha_1+\frac{\beta}{2}}\xi Ue^{s\varphi}\right]\, dxdt \nonumber \\
& \leq C \int_0^T\int_0^1 \left( x^{2\alpha_1-\beta -2}\xi^2U^2 +
x^{2-2\alpha_1+\beta }\xi^2U^2\right)e^{2s\varphi} dxdt \nonumber \\
& \leq C\int_0^T\int_0^1 x^{2\alpha_1-\beta }(\xi Ue^{s\varphi})_x^2 dxdt
+
C\int_0^T\int_0^1 x^{2-2\alpha_1+\beta }\xi^2U^2e^{2s\varphi} dxdt \nonumber \\
& \leq C \int_0^T\int_0^1 \left( x^{2\alpha_1-\beta }\xi^2 U_x^2+x^{2\alpha_1-\beta }\xi_x^2 U^2+s^2\Theta^2x^{2+2\alpha_1-3\beta}\xi^2U^2\right)e^{2s\varphi}dxdt \nonumber \\
&\hspace*{.5cm} + C \int_0^T\int_0^1 x^{2-2\alpha_1+\beta }\xi^2U^2e^{2s\varphi} dxdt.
\end{align*}
So since $\beta\geq \alpha$, $\xi_x$ is supported in $\omega'$ and $\Theta$ is bounded below then for $s$ large enough we have
\begin{align}\label{b11U}
\bar{C} \int_0^T\int_0^1 b_{11}^2 \xi^2U^2e^{2s\varphi}dxdt \, \leq \,\,\, & \frac14 \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_1-\beta} \xi^2 U_x^2+ s^3 \Theta^3(t)x^{2+2\alpha_1-3\beta}
\xi^2U^2]e^{2s\varphi}dx dt
\nonumber \\
& + C\int_0^T\int_{\omega'} U^2e^{2s\varphi}dx dt
\end{align}
Similarly, for $s$ large enough we have
\begin{align}\label{b21V}
\bar{C} \int_0^T\int_0^1 b_{21}^2 \xi^2V^2e^{2s\varphi}dxdt \, \leq \,\,\, & \frac14 \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_2-\beta} \xi^2 V_x^2+ s^3 \Theta^3(t)x^{2+2\alpha_2-3\beta}
\xi^2V^2]e^{2s\varphi}dx dt
\nonumber \\
& + C\int_0^T\int_{\omega'} V^2e^{2s\varphi}dx dt
\end{align}
Combining \eqref{estimU}, \eqref{b11U} and \eqref{b21V} we deduce the estimate
\begin{align}\label{estimU2}
\int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_1-\beta}
\xi^2U_x^2(t,x) & +s^3 \Theta^3(t)x^{2+2\alpha_1-3\beta}
\xi^2U^2(t,x)]e^{2s\varphi}dx dt
\notag \\
& \leq \,\, \frac14 \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_1-\beta} \xi^2 U_x^2+ s^3 \Theta^3(t)x^{2+2\alpha_1-3\beta}
\xi^2U^2]e^{2s\varphi}dx dt
\nonumber \\
& \,\,\,\,\,\,\,+ \frac14 \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_2-\beta} \xi^2 V_x^2+ s^3 \Theta^3(t)x^{2+2\alpha_2-3\beta}
\xi^2V^2]e^{2s\varphi}dx dt
\nonumber \\
& \,\,\,\,\,\,\, + C\int_0^T\int_{\omega'} s\Theta(U^2+V^2)e^{2s\varphi}dx dt.
\end{align}
For the second component, Arguing as before we have for $s$ large enough
\begin{align}\label{estimV2}
\int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_2-\beta}
\xi^2V_x^2(t,x) & +s^3 \Theta^3(t)x^{2+2\alpha_2-3\beta}
\xi^2V^2(t,x)]e^{2s\varphi}dx dt
\notag \\
& \leq \,\, \frac14 \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_2-\beta} \xi^2 V_x^2+ s^3 \Theta^3(t)x^{2+2\alpha_2-3\beta}
\xi^2V^2]e^{2s\varphi}dx dt
\nonumber \\
& \,\,\,\,\,\,\,+ \frac14 \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_1-\beta} \xi^2 U_x^2+ s^3 \Theta^3(t)x^{2+2\alpha_1-3\beta}
\xi^2U^2]e^{2s\varphi}dx dt
\nonumber \\
& \,\,\,\,\,\,\, + C\int_0^T\int_{\omega'} s\Theta(U^2+V^2)e^{2s\varphi}dx dt.
\end{align}
Therefore, from \eqref{estimU2} and \eqref{estimV2} we deduce the estimate
\begin{align}\label{estimUV1}
\int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_1-\beta}
\xi^2U_x^2(t,x) & +s^3 \Theta^3(t)x^{2+2\alpha_1-3\beta}
\xi^2U^2(t,x)]e^{2s\varphi}dx dt
\notag \\
+ \int_0^T\int_0^1 [s\Theta(t)x^{2\alpha_2-\beta}
\xi^2V_x^2(t,x) & +s^3 \Theta^3(t)x^{2+2\alpha_2-3\beta}
\xi^2V^2(t,x)]e^{2s\varphi}dx dt
\notag \\
\leq C\int_0^T\int_{\omega'} s\Theta(U^2+V^2)e^{2s\varphi}dx dt.
\end{align}
This gives an estimate on $(0,a')$.
As above, to obtain an estimate on $(a',1)$, we apply \eqref{estimright} to each equation of the system \eqref{adjfirst}-\eqref{adjend}, we use Hardy-Poincar\'e inequality and we obtain the estimate
\begin{align}\label{estimUV2}
\int_0^T\int_0^1 [s\phi
\zeta^2(U_x^2+V_x^2) +s^3 \phi^3\zeta^2(U^2+V^2)]e^{2s\Phi}dx dt
\notag \\
\leq C\int_0^T\int_{\omega'}s^3\phi^3 (U^2+V^2)e^{2s\Phi}dx dt.
\end{align}
Consequently, using \eqref{estimUV1}, \eqref{estimUV2} and \eqref{compare varphi and Phi} we deduce the global estimate
\begin{align*}
& \int_0^T\int_0^1 s\Theta(t) \left[ x^{2\alpha_1-\beta}U_x^2(t,x)+x^{2\alpha_2-\beta}V_x^2(t,x)\right]e^{2s\varphi}\,dx dt \notag \\
& +\int_0^T\int_0^1 s^3 \Theta^3(t)\left[
x^{2+2\alpha_1-3\beta}U^2(t,x)+x^{2+2\alpha_2-3\beta}V^2(t,x)\right] e^{2s\varphi}\,dx dt \notag \\
& \vspace*{2cm}\leq C\int_0^T\int_{\omega'} s^3
\Theta^3
\left[U^2(t,x)+V^2(t,x)\right]e^{2s\Phi} \,dxdt.
\end{align*}
This ends the proof.
\end{proof}
As above, using density argument we deduce the following result for the general case: $U_0, \, V_0 \in L^2(0,1)$.
\begin{corollary}\label{??????}
Let $T>0$ and $(\alpha_1,\alpha_2)\in (0,1)\times(0,1)$ be given.
Let $\beta \in [max(\alpha_1,\alpha_2),1[$ and $\mu_i \geq max(0,2+2\alpha_i-3\beta)$.
Then, there exist two positive
constants $C$ and $s_0$ such that every solution $(U,V)$ of
\eqref{adjfirst}-\eqref{adjend} satisfies
\begin{align}
& \int_0^T\int_0^1 s\Theta(t) \left[ x^{\alpha_1}U_x^2(t,x)+x^{\alpha_2}V_x^2(t,x)\right]e^{2s\varphi (t,x)}\,dx dt \notag \\
& +\int_0^T\int_0^1 s^3 \Theta^3(t)\left[
x^{\mu_1}U^2(t,x)+x^{\mu_2}V^2(t,x)\right] e^{2s\varphi (t,x)}\,dx dt \notag \\
& \vspace*{2cm}\leq C\int_0^T\int_{\omega'} s^3
\Theta^3(t) \left[U^2(t,x)+V^2(t,x)\right]e^{2s\Phi(t,x)}
\,dxdt \quad \quad \mbox{for all}\,\,s\geq s_0. \label{Carl2degenrL2}
\end{align}
\end{corollary}
%
To study of the null-controllability of the system \eqref{eq:1}-\eqref{eq:2}
we need to show the following Carleman estimate.
\begin{theorem} \label{carl1}
Let $T>0$ be given. Assume moreover that
\begin{align}
b_{21}\geq \mu \,\, \mbox{on} \, [0,T]\times \omega_1 \quad \mbox{for some}\,\, \omega_1\Subset \omega\,\, \mbox{and}\,\, \mu>0. \label{b2supGama}
\end{align}
Then there exist two positive constants $C$ and
$s_0$ such that, every solution $(U,V)$ of
\eqref{adjfirst}-\eqref{adjend} satisfies for all $s\geq s_0$ the estimates
\begin{align}
& \int_0^T\int_0^1 s\Theta(t)\left[x^{2\alpha_1-\beta}U_x^2(t,x)+x^{2\alpha_2-\beta}V_x^2(t,x))\right]e^{2s\varphi (t,x)}\,dx dt\notag \\
& + \int_0^T\int_0^1 s^3 \Theta^3(t)
\left[x^{2+2\alpha_1-3\beta}U^2(t,x)+x^{2+2\alpha_2-3\beta}V^2(t,x)\right]e^{2s\varphi (t,x)}\,dx dt\notag
\\
& \leq C\int_0^T\int_\omega U^2(t,x)\,dx dt. \label{carleman2}
\end{align}
\end{theorem}
\begin{remark}
The assumption \eqref{b2supGama} can be replaced by \\
\hspace*{4,3cm} $b_{21}\leq -\mu$ \,\, on \, $[0,T]\times \omega_1$ \quad for some \, $\omega_1\Subset \omega$\, and\, $\mu>0$.
\end{remark}
Theorem \ref{carl1} is a consequence of Theorem \ref{carl01} applied to $\omega_1$ and the following lemma, see also the proofs of (\cite{Ca-Te}, Theorem 3.2), \cite{Liu} and \cite{bahm}.
\begin{lemma}\label{lemma} Suppose moreover that \eqref{b2supGama} holds. Then
for all $\varepsilon >0$ there exists a positive constant $C_\varepsilon >0$ such that every solution $(U,V)$ of
\eqref{adjfirst}-\eqref{adjend} satisfies
\begin{align}\label{estilemma}\int_0^T\int_{\omega_1} s^3\Theta^3 V^2
e^{2s\Phi}dx dt& \leq \varepsilon J(V)+C_\varepsilon\int_0^T\int_{\omega} U^2 dx dt,
\end{align}
\end{lemma}
where $\omega_1$ is defined in \eqref{b2supGama} and
$$J(V):=\int_0^T\int_0^1 \left[s\Theta(t) x^{2\alpha_2-\beta}V_x^2+
s^3 \Theta^3(t) x^{2+2\alpha_2-3\beta}V^2\right]e^{2s\varphi (t,x)}\,dx dt
$$
\begin{proof}
Let $\chi\in\mathcal{C}^\infty(0,1)$ such that $supp \,\chi \subset
\omega$ and $\chi \equiv 1$ on $\omega_1$. Multiplying the equation
\eqref{adjfirst} by $s^3\Theta^3\chi e^{2s\Phi} V$ and integrating, we obtain
\begin{align}\label{lem1forc1}
\int_0^T\int_0^1 \chi b_{21} s^3\Theta^3 e^{2s\Phi} V^2 dx dt
= & -\int_0^T\int_0^1 \chi s^3\Theta^3 e^{2s\Phi}VU_t\,dxdt + \int_0^T\int_0^1 \chi s^3\Theta^3 e^{2s\Phi}V(x^{\alpha_1}U_x)_x dxdt \notag \\
& - \int_0^T\int_0^1 \chi b_{11}s^3\Theta^3 e^{2s\Phi}UV dxdt
\end{align}
Integrating by parts and using the equation \eqref{21}, we obtain
\begin{align}\label{lem1forc2}
\int_0^T\int_0^1\chi s^3\Theta^3 e^{2s\Phi}VU_t\,dxdt =
& \int_0^T\int_0^1 \chi x^{\alpha_2} s^3\Theta^3 e^{2s\Phi}U_xV_x\,dxdt
+ \int_0^T\int_0^1 x^{\alpha_2} s^3\Theta^3(\chi e^{2s\Phi})_x UV_x \,dxdt \notag \\
& + \int_0^T\int_0^1 \chi b_{12} s^3\Theta^3 e^{2s\Phi}U^2\,dxdt
+ \int_0^T\int_0^1 \chi b_{22} s^3\Theta^3 e^{2s\Phi}UV \,dxdt \notag \\
& -\int_0^T\int_0^1 \chi s^3\left(\Theta^3 e^{2s\Phi}\right)_t UV \,dxdt ,
\end{align}
and
\begin{align}\label{lem1forc3}
& \int_0^T\int_0^1\chi s^3\Theta^3 e^{2s\Phi}V(x^{\alpha_1}U_x)_x\,dxdt =
- \int_0^T\int_0^1 x^{\alpha_1}\chi s^3\Theta^3 e^{2s\Phi}U_xV_x\,dxdt \notag \\
& \hspace*{1cm} + \int_0^T\int_0^1 s^3\Theta^3 x^{\alpha_1}(\chi e^{2s\Phi})_x UV_x \,dxdt
+ \int_0^T\int_0^1 s^3\Theta^3 (x^{\alpha_1}(\chi e^{2s\Phi})_x)_x UV\,dxdt.
\end{align}
So combining the identities \eqref{lem1forc1}-\eqref{lem1forc3}, we get
\begin{align}
& \int_0^T\int_0^1 b_{21} \chi s^3\Theta^3 e^{2s\Phi} V^2 dx dt
\, = \,\, - \,\, \underset{I_1}{\underbrace{ \int_0^T\int_0^1 (x^{\alpha_1}+x^{\alpha_2})\chi s^3\Theta^3 e^{2s\Phi}U_xV_x\,dxdt}} \notag \\
& \hspace*{1.5cm} + \, \underset{I_2}{\underbrace{ \int_0^T\int_0^1 (x^{\alpha_1}-x^{\alpha_2}) s^3\Theta^3(\chi e^{2s\Phi})_x UV_x \,dxdt}} \, - \, \underset{I_3}{\underbrace{ \int_0^T\int_0^1 b_{12}\chi s^3\Theta^3 e^{2s\Phi}U^2\,dxdt}}\notag \\
& \hspace*{1.5cm} +\, \underset{I_4}{\underbrace{ \int_0^T\int_0^1 \left[ s^3\chi\left(\Theta^3 e^{2s\Phi}\right)_t
+s^3\Theta^3(x^{\alpha_1}(\chi e^{2s\Phi})_x)_x
-(b_{11}+b_{22})\chi s^3\Theta^3 e^{2s\Phi}
\right]UV \,dxdt}}. \label{lem1forc4}
\end{align}
Now we estimate the integrals $I_1,\, I_2,\, I_3 $ and $I_4.$
We have
\begin{align}
& \left|\int_0^T\int_0^1 x^{\alpha_i}\chi s^3\Theta^3 e^{2s\Phi}U_xV_x\,dxdt\right| =
\left| \int_0^T\int_0^1 \left[s^{\frac12}\Theta^{\frac12} x^{\alpha_2-\frac{\beta}{2}}e^{s\varphi }V_x\right]\left[s^{\frac52}\Theta^{\frac52} \chi x^{\alpha_i-\alpha_2+\frac{\beta}{2}}e^{s(2\Phi-\varphi)} U_x\right]dx dt\right|\notag \\
& \leq \varepsilon \int_0^T\int_0^1 s\Theta x^{2\alpha_2-\beta}e^{2s\varphi }V^2_x\,dxdt
+ \frac{1}{4\varepsilon} \underset{K}{\underbrace{\int_0^T\int_0^1s^{5}\Theta^{5} \chi^2 x^{2\alpha_i-2\alpha_2+\beta}e^{2s(2\Phi-\varphi)} U_x^2 dx dt}} . \label{lem1forc5}
\end{align}
The last integral $K$ should be estimated by an integral in $U^2$. For this, we multiply the equation \eqref{adjfirst}
by $
s^{5}\Theta^{5}\chi^2x^{\mu}e^{2s(2\Phi-\varphi)}U$ where $\mu := 2\alpha_i-\alpha_1-2\alpha_2+\beta$, we integrate by parts and we obtain
\begin{align*}
K = & \,\, \frac12 \underset{K_1}{\underbrace{\int_0^T\int_0^1s^5\left(\Theta^5e^{2s(2\Phi-\varphi)}\right)_t
\chi^2x^\mu U^2 dx dt}} \\
& +\frac12 \underset{K_2}{\underbrace{\int_0^T\int_0^1 s^5\Theta^5 (x^{\alpha_1}(\chi^2x^\mu e^{2s(2\Phi-\varphi)})_x)_xU^2 \, dxdt}}
-\underset{K_3}{\underbrace{\int_0^T\int_0^1 b_{11}s^5\Theta^5 \chi^2x^\mu e^{2s(2\Phi-\varphi)}U^2 \, dxdt}} \\
& -\underset{K_4}{\underbrace{\int_0^T\int_0^1 b_{21} s^5\Theta^5 \chi^2x^\mu e^{2s(2\Phi-\varphi)}UV \, dxdt}}.
\end{align*}
Since $|\Theta^{'}|\leq C\Theta^2$ and $supp \chi \subset \omega$
we have for $i\in \{1,2,3\}$
\begin{align*}
& |K_i| \leq C\int_0^T\int_\omega s^7\Theta^7 e^{2s(2\Phi-\varphi)} U^2 dx dt,
\end{align*}
For $i=4$ we have
\begin{align*}
|K_4| & =
\int_0^T\int_0^1 [s^{\frac32}\Theta^{\frac32}x^{1+\alpha_2-\frac{3\beta}{2}} e^{s\varphi }V]
[s^{\frac{7}{2}}\Theta^{\frac{7}{2}}b_{21}\chi^2x^{\mu-1-\alpha_2+\frac{3\beta}{2}}
e^{s(4\Phi-3\varphi)}U ]dx dt \notag \\
& \leq \varepsilon^2 \int_0^T\int_0^1
s^3\Theta^3 x^{2+2\alpha_2-3\beta} e^{2s\varphi }V^2 dxdt
+ C_\varepsilon \int_0^T\int_\omega
s^7\Theta^7
e^{2s(4\Phi-3\varphi)} U^2 dx dt.
\end{align*}
So, thanks to \eqref{comparaison phi} we have
\begin{align}
|K|\leq & \,\,\varepsilon^2 \int_0^T\int_0^1 s^3\Theta^3 x^{2+2\alpha_2-3\beta} e^{2s\varphi }V^2 dxdt
+ C_\varepsilon\int_0^T\int_\omega U^2 dx dt. \label{lem1forc11}
\end{align}
From \eqref{lem1forc5}-\eqref{lem1forc11} we deduce the estimate
\begin{align}
|I_1|\leq \,\, 2\varepsilon \int_0^T\int_0^1 s\Theta x^{2\alpha_2-\beta}e^{2s\varphi }V^2_x\,dxdt
+ \frac\varepsilon2 \int_0^T\int_0^1 s^3\Theta^3 x^{2+2\alpha_2-3\beta} e^{2s\varphi }V^2 dxdt
+ C_\varepsilon \int_0^T\int_\omega U^2 dx dt. \label{lem1forc12}
\end{align}
Similarly we have
\begin{align}
|I_2| \leq & \,\, C \int_0^T\int_{\omega} s^4\Theta^4 |UV_x|e^{2s\Phi} dxdt \nonumber
\\
\leq & \,\, \varepsilon \int_0^T\int_0^1 s\Theta x^{2\alpha_2-\beta}e^{2s\varphi }V^2_x\,dxdt +
C_\varepsilon \int_0^T\int_\omega U^2 dx dt, \label{lem1forc13} \\
|I_3| \leq & \,\, C \int_0^T\int_\omega U^2 dx dt, \label{lem1forc15}\\
|I_4| \leq & \,\, C \int_0^T\int_{\omega} s^6\Theta^6 |UV|e^{2s\Phi} dxdt \nonumber
\\
\leq & \,\, \varepsilon \int_0^T\int_0^1 s^3\Theta^3 x^{2+2\alpha_2-3\beta}e^{2s\varphi }V^2\,dxdt +
C_{\varepsilon} \int_0^T\int_\omega U^2 dx dt. \label{lem1forc14}
\end{align}
Consequently, from the estimates \eqref{lem1forc12}-\eqref{lem1forc14}, we conclude that
\begin{align*}
\int_0^T\int_0^1 b_{21} \chi e^{2s\varphi } V^2 dx dt \leq 3\varepsilon J(V) + C_{\varepsilon} \int_0^T\int_\omega U^2 dx dt.
\end{align*}
Finally, since $\chi \equiv 1$ on $\omega_1$, then using \eqref{b2supGama} we achieve the claim.
\end{proof}
As above, using a density argument we deduce the following result for the general case: $U_0,\, V_0\in L^2(0,1)$.
\begin{corollary} \label{corol1forc}
Let $T>0$ be given. Assume moreover that \eqref{b2supGama} holds.
Let $(\alpha_1,\alpha_2)\in (0,1)\times(0,1)$, $\beta \in [max(\alpha_1,\alpha_2),1[$ and $\mu_i \geq max(0,2+2\alpha_i-3\beta)$.
Then, there exist two positive constants $C$ and
$s_0$ such that, every solution $(U,V)$ of
\eqref{adjfirst}-\eqref{adjend} satisfies, for all $s\geq s_0$ the estimates
\begin{align}
& \int_0^T\int_0^1 s\Theta(t)\left[x^{\alpha_1}U_x^2(t,x)+x^{\alpha_2}V_x^2(t,x))\right]e^{2s\varphi (t,x)}\,dx dt\notag \\
& + \int_0^T\int_0^1 s^3 \Theta^3(t)
\left[x^{\mu_1}U^2(t,x)+x^{\mu_2}V^2(t,x))\right]e^{2s\varphi (t,x)}\,dx dt\notag
\\
& \leq C\int_0^T\int_\omega U^2(t,x)\,dx dt. \label{carlemanonef}
\end{align}
\end{corollary}
\section{Observability and null controllability of linear systems}
As a consequence of the Carleman
estimates established in the above section, we prove first a observability
inequality for the adjoint problem \eqref{adjfirst}-\eqref{adjend}
of problem \eqref{eq:1}-\eqref{eq:2}.
\begin{theorem}\label{observatos}
Let $T>0$ be given. Assume that \eqref{b2supGama} is satisfied. Then, there exists a positive constant $C$ such
that every solution $(U,V)$ of \eqref{adjfirst}-\eqref{adjend}
satisfies
\begin{align}
\int_0^1\left[U^2(T,x)+V^2(T,x)\right]dx\leq
C\int_0^T\int_{\omega}U^2(t,x)dx dt. \label{observ1forc}
\end{align}
\end{theorem}
\begin{proof}
Multiplying the equations \eqref{adjfirst} and \eqref{21}
respectively by $U_t$ and $V_t$ and integrating over $(0,1)$ the sum of the new equations we obtain
\begin{align*}
0=&\int_0^1\left[U_t^2+V_t^2\right]dx -\left[x^{\alpha_1}
U_xU_t\right]_{x=0}^1-\left[x^{\alpha_2} V_xV_t\right]_{x=0}^1
\\
&+\int_0^1 b_{11}U U_t\,dx +\int_0^1 b_{22}V V_t+\int_0^1 b_{21}V U_t\,dx\\
&+\int_0^1 b_{12}U V_t\,dx+\frac{1}{2}\frac{d}{dt}\int_0^1[x^{\alpha_1}U_x^2+x^{\alpha_2}V_x^2]\,dx.
\end{align*}
Using the Young's inequality we obtain
\begin{align*}
\frac{1}{2}\frac{d}{dt}\int_0^1[x^{\alpha_1}U_x^2+x^{\alpha_2}V_x^2]\,dx &\leq \int_0^1(b_{11}^2+b_{12}^2)U^2dx +\int_0^1(b_{22}^2+b_{21}^2)V^2dx\\
&\leq C\int_0^1(U^2(t,x)+V^2(t,x))\,dx\\
&\leq C\int_0^1[x^{\alpha_1 -2}U^2(t,x)+x^{\alpha_2 -2}V^2(t,x)]dx.
\end{align*}
Hence, using the Hardy-Poincar\'e inequality \eqref{Hardy} one has
\begin{align*}
\frac{d}{dt}\int_0^1[x^{\alpha_1}U_x^2+x^{\alpha_2}V_x^2]\,dx \leq C_0 \int_0^1[x^{\alpha_1}U_x^2+x^{\alpha_2}V_x^2]\,dx
\end{align*}
Hence
\begin{align*}
\frac{d}{dt}\left\{e^{-C_0t}\int_0^1[x^{\alpha_1}U_x^2+x^{\alpha_2}V_x^2]dx\right\}&\leq 0.
\end{align*}
Consequently, the function $t\longmapsto
e^{-C_0t}\int_0^1[x^{\alpha_1}U_x^2+x^{\alpha_2}V_x^2]dx $ is not increasing. Thus,
$$\int_0^1[x^{\alpha_1}U_x^2(T,x)+x^{\alpha_2}V_x^2(T,x)]dx\leq
e^{C_0T}\int_0^1[x^{\alpha_1}U_x^2(t,x)+x^{\alpha_2}V_x^2(t,x)]dx.
$$
Integrating over
$[\frac{T}{4},\frac{3T}{4}]$ and using the Carleman estimate \eqref{carlemanonef} one obtains
\begin{align*}
\int_0^1[x^{\alpha_1}U_x^2(T,x)+x^{\alpha_2}V_x^2(T,x)]dx&\leq \frac{2e^{C_0T}}{T}\int_{\frac{T}{4}}^{\frac{3T}{4}}\int_0^1[x^{\alpha_1}U_x^2(t,x)+x^{\alpha_2}V_x^2(t,x)]dx dt
\nonumber
\\ &\leq C_T \int_{\frac{T}{4}}^{\frac{3T}{4}}\int_0^1s\Theta e^{2s\varphi}[x^{\alpha_1}U_x^2(t,x)+x^{\alpha_2}V_x^2(t,x)]dx dt
\nonumber
\\
&\leq C_T\int_0^T\int_\omega U^2(t,x)
dx dt,
\end{align*}
On the other hand, using hardy-Poincar\'e inequality one gets
\begin{align*}
\int_0^1[U^2(T,x)+V^2(T,x)]dx &\leq \int_0^1[x^{\alpha_1-2}U^2(T,x)+x^{\alpha_2-2}V^2(T,x)]dx
\\
& \leq C \int_0^1[x^{\alpha_1}U_x^2(T,x)+x^{\alpha_2}V_x^2(T,x)]dx
\nonumber
\end{align*}
This ends the proof.
\end{proof}
By Theorem \ref{observatos} and a classical argument one can deduce the controllability result
\begin{theorem}
If
the assumption \eqref{b2supGama} is satisfied, then the degenerate coupled system \eqref{eq:1}-\eqref{eq:2}
is null controllable.
\end{theorem}
\section{Appendix 1}
As in \cite{bouss}, \cite{Ca-Te}, \cite{bahm}, we give the proof of the Caccioppoli's inequality for degenerate coupled systems with two different diffusion coefficients.
\begin{lemma}\label{cacci}
Let $\omega' \Subset\omega$. Then there exists a positive constant C such that
\begin{align*}
\int_0^T\int_{\omega'} \left[
U_x^2(t,x)+V_x^2(t,x)\right]e^{2s\varphi_i}dx dt \leq
C\int_0^T\int_{\omega} \left[ U^2(t,x)+V^2(t,x)\right]e^{2s\varphi_i}dx dt.
\end{align*}
\end{lemma}
\begin{proof}
Let $\chi\in\mathcal{C}^\infty(0,1)$ such that $supp\,\chi \subset
\omega$ and $\chi \equiv 1$ on $\omega'.$ We have
\begin{align*}
0=&\int_0^T\frac{d}{dt} \left[ \int_0^1\chi^2
(U^2+V^2) e^{2s\varphi_i}dx \right] dt\\
=&
-2\int_0^T\int_0^1 \chi^2 x^{\alpha_1}U_x^2 e^{2s\varphi_i}dx dt
-2\int_0^T\int_0^1 \chi^2 x^{\alpha_2}V_x^2 e^{2s\varphi_i}dx
\\
&+ \int_0^T\int_0^1 (x^{\alpha_1}(\chi^2 e^{2s\varphi_i})_x)_x U^2 dx +
\int_0^T\int_0^1 (x^{\alpha_2}(\chi^2 e^{2s\varphi_i})_x)_x V^2 dx
\\
& -2\int_0^T\int_0^1 b_{11}\chi^2 U^2 e^{2s\varphi_i} dxdt
-2\int_0^T\int_0^1 b_{22}\chi^2 V^2e^{2s\varphi_i}dxdt
\\
& 2\int_0^T\int_0^1 s \dot{\varphi_i}\chi^2(U^2+V^2)e^{2s\varphi_i}dxdt
-2\int_0^T\int_0^1 (b_{12}+b_{21})\chi^2 U V e^{2s\varphi_i}dx dt
\end{align*}
Therefore, since $\chi$ is supported in $\omega$ and $\chi\equiv 1$ in $\omega'$ then, using Young inequality one obtains
\begin{align*}
\int_0^T\int_{\omega'} (U_x^2+ V_x^2) e^{2s\varphi_i}dxdt
& \leq C \int_0^T\int_0^1 \chi^2 (x^{\alpha_1}U_x^2+x^{\alpha_2} V_x^2) e^{2s\varphi_i}dxdt
\\
&\leq C \int_0^T\int_\omega
(U^2+V^2)e^{2s\varphi_i}dxdt.
\end{align*}
This ends the proof.
\end{proof}
\section{Appendix 2}
\begin{lemma}\label{Hardy2}
For all $\gamma <1$ and all $v$ locally absolutely continuous on $(0,1]$, continuous at $0$ and satisfying $v(0)=0$ and $\int_0^1x^\gamma v_x^2dx <+\infty$ the following Hardy-Poincar\'e inequality holds
\begin{align}\label{Hardy}
\int_0^1x^{\gamma-2}v^2dx\leq C_\gamma \int_0^1x^\gamma v_x^2dx \quad\quad \mbox{where}\,\,\,\,
C_\gamma=\frac{4}{(1-\gamma)^2}
\end{align}
\end{lemma}
\begin{proof}
This result was proved by Cannarsa et al. in \cite{bouss} for $\gamma \in (0,1)$, but by a careful examination of the proof one can see that it remains valid for all $\gamma <1$. In fact let $\gamma <1$ and $\delta= \frac{\gamma+1}{2}$. Using Holder inequality and Fubini's theorem one has
\begin{align}
\int_0^1x^{\gamma-2}v^2dx & = \int_0^1x^{\gamma-2}\left(\int_0^x y^{\delta/2}v'(y) y^{-\delta/2}dy\right)^2 dx \nonumber \\
& \leq \int_0^1x^{\gamma-2}\left(\int_0^x y^{\delta}|v'(y)|^2dy\right)\left( \int_0^x y^{-\delta}dy\right) dx \nonumber \\
&= \frac{1}{1-\delta} \int_0^1 \int_0^x x^{\gamma-\delta-1} y^{\delta}|v'(y)|^2dy dx \nonumber \\
&= \frac{1}{1-\delta} \int_0^1
\left(\int_y^1 x^{\gamma-\delta-1}dx \right)
y^{\delta}|v'(y)|^2 dy \nonumber \\
& \leq \frac{1}{(1-\delta)(\delta-\gamma)} \int_0^1 y^{\gamma}|v'(y)|^2 dy \nonumber \\
& = \frac{4}{(1-\gamma)^2} \int_0^1 y^{\gamma}|v'(y)|^2 dy. \nonumber
\end{align}
This ends the proof.
\end{proof}
\section{Conclusion}
In this paper, we studied the null controllability of linear degenerate systems with two different coefficients diffusion not necessarily of the cascade form. We developed new Carleman estimates. By a standard linearization argument and fixed point, see \cite{bahm}, \cite{bouss}, \cite{Can-Gen-2006}, \cite{Zuazua99}, one can show easily the null controllability of semilinear degenerate coupled systems with two different diffusion coefficients. In this paper we studied coupled system of two weakly degenerate equations. The cases when one of the equation is strongly degenerate systems are open.
|
\section{Introduction}
Finitely generated shift-invariant spaces have been widely used in approximation theory, numerical analysis, sampling theory and
wavelet theory ( see e.g., \cite{ACHM07,AG01, AST05, Bow00, CS07, DDR94, RS95, UB00} and the references therein).
Shift-invariant spaces with additional invariance have been studied in the context of
wavelet analysis and sampling theory \cite{AKTW11, CS03,HL09, Web00}, and have been given a complete algebraic description in \cite {ACHKM10} for $L^2(\hbox{\ensuremath{\mathbb{R}}})$ and in \cite {ACP10} for $L^2(\hbox{\ensuremath{\mathbb{R}}}^d)$. As a tool for showing our main results, we will prove a slightly different but useful characterization.
It is well-known that the Paley-Wiener space $PW$ is translation-invariant. Moreover, Shannon's sampling theorem easily implies that $PW$ is principal, i.e. generated by the single function $\text{sinc}$. It turns out that the fact that $\text{sinc}$ is non-integrable is not a coincidence. Actually, for a principal shift-invariant space in $L^2(\hbox{\ensuremath{\mathbb{R}}})$ which is translation-invariant, any frame generator is non-integrable (see for instance \cite{ASW11}). This observation holds in any dimension. Indeed, the Fourier transform $\widehat{\phi}$ of a frame generator has to satisfy for a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}^d$
$$C^{-1}1_A(\omega)\leq |\widehat{\phi}(\omega)|\leq C 1_{A}(\omega) $$
for some $C\geq 1$ and some finite measure subset $A$. In particular, $\widehat{\phi}$ is not continuous. Such a condition also prevents $\widehat{\phi}$ from being in the Sobolev space\footnote{i.e. $\int|\phi(x)|^2(1+|x|)dx=\infty$. } $H^{\frac 12}(\hbox{\ensuremath{\mathbb{R}}}^d)$ (see \cite{KW99}), whereas it belongs to $H^{\frac 12-\epsilon}(\hbox{\ensuremath{\mathbb{R}}}^d)$ for every $\epsilon>0$ when $A$ is a Euclidean ball of positive radius.
Our first result is a straightforward generalisation of this fact to shift-invariant spaces generated by several functions.
\begin{theorem}\label{notranslation}
Let $\Lambda$ be a lattice in $\hbox{\ensuremath{\mathbb{R}}}^d$.
If a finitely generated $\Lambda$-invariant space of $L^2(\hbox{\ensuremath{\mathbb{R}}}^d)$ is translation-invariant, then at least one of its frame generators has a non continuous Fourier transform, and in particular is not in $L^1$. Moreover this generator satisfies $\int|\phi(x)|^2(1+|x|)dx=\infty$.
\end{theorem}
The slow spatial-decay of the generators of shift-invariant spaces that are also translation-invariant is a disadvantage for the numerical implementation of some analysis and processing algorithms. This is a motivation for considering instead shift-invariant spaces that are only $\frac 1n\hbox{\ensuremath{\mathbb{Z}}}$-invariant and hoping the generators will have better time-frequency localization. Indeed, it was shown in \cite{ASW11} that for every $n$, one can construct a principal shift-invariant space with an orthonormal generator which is in $L^1(\hbox{\ensuremath{\mathbb{R}}})$ although the space is $\frac 1n\hbox{\ensuremath{\mathbb{Z}}}$-invariant.
However, there are still obstructions if we require more regularity on the Fourier transform of the generators. Asking for fractional differentiability yields the following Balian-Low type obstructions (see \cite{BCGP03} and the reference therein). Compare \cite[Theorem 1.2]{ASW11}.
\begin{theorem}\label{epsilonthm}
Let $\Lambda<\Gamma$ be two lattices\footnote{The notation $\Lambda<\Gamma$ is standard to denote subgroups.} in $\hbox{\ensuremath{\mathbb{R}}}^d$.
Suppose $\phi_i\in L^2(\hbox{\ensuremath{\mathbb{R}}}^d)$ are such that $\{\phi_i(\cdot+\lambda)| \lambda\in \Lambda, i=1, \dots, r\}$ forms a frame for the closed subspace $V^{\Lambda}(\Phi)$ spanned by these functions. Let $\rho$ ($\leq r$) be the minimal number of generators of $V^{\Lambda}(\Phi)$. Assume that $[\Gamma:\Lambda]$ is not a divisor of $\rho$, and suppose that $V^{\Lambda}(\Phi)$ is $\Gamma$-invariant. Then there exists $i_0\in\{1, \dots, r\}$ such that
$$\int_{\hbox{\ensuremath{\mathbb{R}}}^d} |\phi_{i_0}(x)|^2 |x|^{d+\epsilon} dx=+\infty$$
for all $\epsilon>0$. I.e. $\widehat{\phi}_{i_0}$ is not in $H^{\frac d2+\epsilon}(\hbox{\ensuremath{\mathbb{R}}}^d)$.
\end{theorem}
One can also ask for a combinaison of regularity, namely continuity and of control of the decay at infinity. In this spirit, we obtain the following result (compare \cite[Theorem 1.3]{ASW11}).
\begin{theorem}\label{pointwise}
Keep the same assumptions as in the last theorem, and suppose moreover that $\widehat{\phi}_i$ is continuous for every $i=1, \dots, r$. Then there exits $i_0\in\{1, \dots, r\}$ such that $\omega^{\frac d2+\epsilon} \widehat\phi_{i_0}(\omega)\not\in L^\infty(\hbox{\ensuremath{\mathbb{R}}}^d)$ for any $\epsilon>0$, i.e.,
\begin{equation*}
{\rm ess\,sup}\ _{\omega\in\mathbb R} |\widehat\phi_{i_0}(\omega)| |\omega|^{\frac d2+\epsilon} =+\infty.
\end{equation*}
\end{theorem}
For $d=1$, the exponent is sharp up to the $\epsilon$ in both theorems but it does not seem to be the case for larger $d$. We actually conjecture that the right exponent should be the same for all dimensions.
Notice that the condition $[\Gamma:\Lambda]\nmid \rho$ in the above theorems is essential.
Indeed all regularity constraints trivially disappear when $\rho=k[\Gamma:\Lambda]$ for any integer $k>0$. To see why, start with some $\Gamma$-invariant space generated by exactly $k$ orthogonal generators $\phi_1,\ldots, \phi_k$, with --say-- smooth and compactly generated Fourier transforms. Then note that as a $\Lambda$-invariant space, $V^{\Gamma}(\phi_1,\ldots, \phi_k)$ is generated by the orthogonal generators $\phi_i(\cdot-f)$ for $f\in F$ and $i=1,\ldots, k$, where $F$ is a section of $\Gamma/\Lambda$ in $\Gamma$.
The previous results state that under additional invariance there always exists at least one (frame) generator whose Fourier transform has poor regularity. One may wonder if at least some generators can be chosen with good properties. We do not know what the optimal proportion of good generators should be. The following proposition gives a lower bound on the number of good generators. Observe that this bound gets worse with the dimension.
\begin{proposition}\label{prop:regular}
For every $d\geq 1$, and every $k\in \hbox{\ensuremath{\mathbb{N}}}$ there exists an SIS $V(\Phi)$ in $L^2(\hbox{\ensuremath{\mathbb{R}}}^d)$ generated by an orthonormal basis $\Phi$ consisting of $r=(2k)^d$ functions, $k^d$ of which have smooth and compactly supported Fourier transforms, and such that $V(\Phi)$ is translation-invariant. Moreover all the generators can be chosen so that their Fourier transforms are in $H^{\frac 12-\epsilon}(\hbox{\ensuremath{\mathbb{R}}}^d)$ for all $\epsilon>0$.
\end{proposition}
We also have the following result,
\begin{proposition}\label{prop:pointwise}
For $d\geq 1$, let $\Gamma$($>\hbox{\ensuremath{\mathbb{Z}}}^d$) be a lattice of $\hbox{\ensuremath{\mathbb{R}}}^d$ and $r\ge1$. Then there exists an SIS $V(\Phi)$ in $L^2(\hbox{\ensuremath{\mathbb{R}}}^d)$ generated by $r$ orthonormal generators $\phi_i$'s, all of which are in $L^1$ (hence they have continuous Fourier transforms) and satisfy $\omega^{\frac 12} \widehat\phi_i(\omega)\in L^\infty(\hbox{\ensuremath{\mathbb{R}}}^d)$. Moreover, $V(\Phi)$ is $\Gamma$-invariant.
\end{proposition}
To summarize, while Proposition \ref{prop:regular} states that it is possible to construct translation-invariant SIS with a portion of the generators having smooth and compactly supported Fourier transforms, Proposition \ref{prop:pointwise} shows that we can construct $\Gamma$-invariant SIS with all its generators having certain pointwise decay in Fourier domain.
\subsection*{Organization}
In the following section, we state and prove a convenient characterization (similar to the one given in \cite{ACHKM10}) of SIS with additional invariance.
In Section \ref{Section:Frame}, we refine this characterization under the assumption that the generators and their translates form a frame.
In Section \ref{section:continuous}, we prove a useful property of the Gramian under the sole assumption that the generators have continuous Fourier transform.
Sections \ref{section:mainresults} and \ref{section:mainpropositions} are dedicated to the proof of the results stated in the introduction.
\section{Finitely generated shift-invariant spaces with additional invariance}
Given a closed subgroup $\Lambda$ of $\hbox{\ensuremath{\mathbb{R}}}^d$, we define its Fourier transform $\Lambda^*$ as the closed subgroup of $\hbox{\ensuremath{\mathbb{R}}}^d$ defined by $$\Lambda^*=\{x\in \hbox{\ensuremath{\mathbb{R}}}^d, e^{2\pi i\langle \lambda ,x\rangle}=1 \; \forall \lambda\in \Lambda\}.$$
Note that the map $\Lambda\to \Lambda^*$ is an involution. The Fourier transform reverses the inclusions, namely if $\Lambda<\Gamma$ then $\Gamma^*< \Lambda^*$.
Observe for instance that for $d=1$, the Fourier transform of $n\hbox{\ensuremath{\mathbb{Z}}}$ is $\frac{1}{n}\hbox{\ensuremath{\mathbb{Z}}}$, or that for $d=2$, the Fourier transform of $\hbox{\ensuremath{\mathbb{R}}}\times \{0\}$ is $\{0\}\times \hbox{\ensuremath{\mathbb{R}}}$.
In this paper, we consider the following general setting: $\Lambda <\Gamma$ are two closed subgroups of $\hbox{\ensuremath{\mathbb{R}}}^n$.
Let $\phi_1,\ldots, \phi_r\in L^2(\hbox{\ensuremath{\mathbb{R}}}^d)$, and denote by $\Phi$ the column vector whose components are the $\phi_i$'s. We will denote by $V^{\Lambda}(\Phi)$ the smallest closed $\Lambda$-invariant subspace containing the $\phi_i$'s.
It is common to call $V^{\Lambda}(\Phi)$ {\it shift-invariant} when $\Lambda=\hbox{\ensuremath{\mathbb{Z}}}^d$, and {\it translation-invariant} when $\Lambda=\hbox{\ensuremath{\mathbb{R}}}^d$. To allege notation, we will omit the subscript $\hbox{\ensuremath{\mathbb{Z}}}^d$, when $\Lambda=\hbox{\ensuremath{\mathbb{Z}}}^d$.
We will be interested in the situation where $V^{\Lambda}(\Phi)$ is in addition $\Gamma$-invariant. As we will see later, this prevents frame generators from having nice decay at infinity.
We will start by providing a short and self-contained proof of a result essentially due to \cite{ACHKM10, ACP10}. Observe that this problem is only non-trivial when the quotient $\hbox{\ensuremath{\mathbb{R}}}^d/\Lambda$ is compact since otherwise, the only finitely generated $\Lambda$-invariant space is $\{0\}$. This is equivalent to the fact that $\Lambda^*$ (hence $\Gamma^*$) is discrete.
In this paper we will mainly focus on the cases when $\Lambda$ is a lattice and when $\Gamma$ is either a (larger) lattice or all of $\hbox{\ensuremath{\mathbb{R}}}^d$.
Before stating this result, let us introduce some notation. The Gramian associated to $\Phi$ and $\Lambda$ is a measurable field of $r\times r$ matrices whose general coefficient is defined for a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}$ by
$$G^{\Lambda}_{i,j}(\omega)= \sum_{l \in \Lambda^*} \widehat{\phi_i}(\omega+l)\overline{\widehat{\phi_j}}(\omega+l).$$
For short, $$G^{\Lambda}(\omega)= \sum_{l\in \Lambda^*}\widehat{\Phi}(\omega+l)\widehat{\Phi}^*(\omega+l).$$
For a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}$, let
$A(\omega)$ to be the $r\times r$ matrix defined by
$$A(\omega)= \sum_{g\in \Gamma^*}\widehat{\Phi}(\omega+g)\widehat{\Phi}^*(\omega+g).$$
Now, let $F$ be a subset of $\Lambda^*$ consisting of representatives of the quotient $\Lambda^*/ \Gamma^*$. For instance, for $d=1$, $\Lambda=\hbox{\ensuremath{\mathbb{Z}}}$ and $\Gamma=\frac{1}{n}\hbox{\ensuremath{\mathbb{Z}}}$, one has $\Lambda^*=\hbox{\ensuremath{\mathbb{Z}}}$ and $\Gamma^*=n\hbox{\ensuremath{\mathbb{Z}}}$, and for $F$, one can take $\{0,\ldots n-1\}$.
We have
\begin{equation}\label{eq:GA}
G^{\Lambda}(\omega)=\sum_{f\in F}A(\omega+f).
\end{equation}
Now let us state the main result of this section. Although it could easily be deduced from the main results in \cite{ACHKM10, ACP10}, we chose to write down a short self-contained proof.
\begin{theorem}\label{thm:rank}
Let $\Lambda<\Gamma$ be closed cocompact subgroups of $\hbox{\ensuremath{\mathbb{R}}}^d$.
The space $V^{\Lambda}(\Phi)$ is $\Gamma$-invariant if and only if the following equality holds for a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}^d$.
\begin{equation}\label{eq:rank}
\mathrm{rank\ } G^{\Lambda}(\omega)=\sum_{f\in F} \mathrm{rank\ }A(\omega+f).
\end{equation}
\end{theorem}
\begin{proof}
Recall that a function $\varphi$ belongs to $V^{\Lambda}(\Phi)$ if and only if there exist bounded measurable $\Lambda^*$-periodic functions $P_1,\ldots,P_r$ such that $\widehat{\varphi}=P_1\widehat{\phi_1}+\ldots+P_r\widehat{\phi_r}$. Fiberwise, this is equivalent to saying that for a.e. $\omega$, the vector $(\widehat{\varphi}(\omega+l))_{l\in \Lambda^*}\in \ell^2(\Lambda^*)$ lies in the subspace $\widehat{V}(\omega)$ spanned by the $r$ vectors $(\widehat{\phi_i}(\omega+l))_{l\in \Lambda^*}$ for $i=1,\ldots ,r$.
Now the space $V^{\Lambda}(\Phi)$ is $\Gamma$-invariant if for every bounded $\Gamma^*$-periodic function $Q$, and for every $i=1,\ldots ,r$, there are bounded measurable $\Lambda^*$-periodic functions $P_{i,1},\ldots,P_{i,r}$ such that $$Q\widehat{\phi_i}=P_{i,1}\widehat{\phi_1}+\ldots+P_{i,r}\widehat{\phi_r}.$$
Again, this is equivalent to saying that for a.e. $\omega$, and for every bounded $\Gamma^*$-periodic $\theta: \; \Lambda^*\to \hbox{\ensuremath{\mathbb{C}}}$ the vector $(\theta(\omega+l)\widehat{\phi_i}(\omega+l))_{l\in \Lambda^*}\in \ell^2(\Lambda^*)$ lies in the subspace $\widehat{V}(\omega)$.
Let us denote $\widehat{W}(\omega)$ the subspace of $ \ell^2(\Lambda^*)$ spanned by $(\theta(\omega+l)\widehat{\phi_i}(\omega+l))_{l\in \Lambda^*}$ for all bounded $\Gamma^*$-periodic $\theta$, and every $i=1,\ldots, r$. Clearly $\widehat{V}(\omega)$ is a subspace of $\widehat{W}(\omega)$, and the two coincide for a.e. $\omega$ exactly when $V^{\Lambda}(\Phi)$ is $\Gamma$-invariant.
The rest of the proof amounts to showing that the left-hand term in (\ref{eq:rank}) corresponds to the dimension of $\widehat{V}(\omega)$ (which is obvious) and that the right-hand term of the equality corresponds to the dimension of $\widehat{W}(\omega)$.
A basis for the space of bounded $\Gamma^*$-periodic functions on $\Lambda^*$ consists of the functions $(\theta_f)_{f\in F}$, where each $\theta_f(l)$ equals $1$ if $l\in f+\Gamma^*$ and $0$ elsewhere. Observe that for all $f\neq f'$, and all $\varphi, \varphi'\in W_{\widehat{\Phi}}(\omega)$, $\theta_f\varphi$ and $\theta_{f'}\varphi'$ are orthogonal.
Hence $\widehat{W}(\omega)$ decomposes as a direct sum $$\widehat{W}(\omega)=\bigoplus_{f\in F}\widehat{W}^f(\omega),$$
where $\widehat{W}^f(\omega)$ is the subspace of functions of the form $\theta_f\varphi$, for $\varphi\in \widehat{W}(\omega).$
But it comes out that $\widehat{W}^f(\omega)$ is precisely the subspace of $\ell^2(\Lambda^*)$ spanned by $(\widehat{\phi_i}(\omega+f+g))_{g\in \Gamma^*}$, whose dimension equals the rank of $A(\omega+f)$. This finishes the proof of the theorem.
\end{proof}
Although we will be mostly interested in the case where both $\Lambda$ and $\Gamma$ are lattices, we will also use the following special case of Theorem \ref{thm:rank}, where $\Lambda$ is a lattice and $\Gamma=\hbox{\ensuremath{\mathbb{R}}}^d$ (so $\Gamma^*=\{0\}$).
\begin{corollary}\label{cor:rank}
Let $\Lambda$ be a lattice, and let $V^{\Lambda}(\Phi)$ be a $\Lambda$-invariant space generated by $\phi_1,\ldots ,\phi_r$. Then it is translation-invariant if and only if the following equality holds for a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}^d$.
\begin{eqnarray}\label{eq:rank2}
\mathrm{rank\ } G^{\Lambda}(\omega) &=&\sum_{f\in \Lambda^*} \mathrm{rank\ }A(\omega+f)\\
& = & \left|\{f\in \Lambda^*, \Phi(\omega+f)\neq 0 \}\right|.
\end{eqnarray}
\end{corollary}
\section{Frame generators and additional invariance}\label{Section:Frame}
The main goal of this section is to reformulate Theorem \ref{thm:rank} under the additional assumption that the generators form a frame for the space $V^{\Lambda}(\Phi)$.
Recall that the family of all $\Lambda$-translates of the generators $\phi_1, \ldots, \phi_r$ form a Riesz basis if and only if there exists $s\geq 1$ such that
\begin{equation}\label{riesz.def}
s^{-1}I\le G^{\Lambda}(\omega)\le sI, \ \text{a.e.} \ \omega\in\hbox{\ensuremath{\mathbb{R}}}.
\end{equation}
In this case the functions $\phi_i$'s are called \emph{Riesz generators} for $V^{\Lambda}(\Phi)$.
Similarly, $\{\phi_i(\cdot+\lambda)|\ \lambda\in \Lambda, i=1, \dots, r\}$ is a frame if and only there exists $s\geq 1$ such that
\begin{equation}\label{upperlower}
s^{-1}G^{\Lambda}(\omega)\le (G^{\Lambda}(\omega))^2\le sG^{\Lambda}(\omega)
\end{equation}
for almost every $\omega\in\hbox{\ensuremath{\mathbb{R}}}^d$ (see \cite{Bow00}). Here the $\phi_i$'s are called {\it frame generators} for $V^{\Lambda}(\Phi)$.
Let us start by an easy lemma.
Given a non-negative self-adjoint matrix $A$, denote $q_A$ its associated quadratic form, $\mu^-(A)$ its smallest non-zero eigenvalue and $k_A$ the dimension of its kernel. Denote the unit sphere of a Euclidean space $V$ by $S_V$.
\begin{lemma}\label{lem:matrices}
Let $C=A+B$ be three $d\times d$ non-negative self-adjoint matrices such that $\mathrm{rank\ } C=\mathrm{rank\ } A+\mathrm{rank\ } B$.
Then $\mu^-(C)\leq \min\{\mu^-(A),\mu^-(B)\}$.
\end{lemma}
\begin{proof}
Note that we can assume without loss of generality that $C$ has full rank. Observe that $\text{Ker} A$ and $\text{Ker} B$ are in direct sum. By the min-max theorem,
$$\mu^-(A)=\min_{\dim V=k_A+1}\;\max_{x\in S_V}q_A(x).$$
Now let $V_0$ be a subspace minimizing the above expression. Let $x\in V_0\cap \text{Ker} B$ of norm $1$. We have
$$\mu^-(A)\geq q_A(x)=q_A(x)+q_B(x)=q_C(x)\geq \mu^-(C).$$
We conclude since the roles of $A$ and $B$ are symmetric.
\end{proof}
The following theorem will play a central role in the sequel.
\begin{theorem}\label{thm:frame}
Let $\Lambda<\Gamma$ be closed cocompact subgroups of $\hbox{\ensuremath{\mathbb{R}}}^d$. Let $\Phi$ be finite set of generators of $V^{\Lambda}(\Phi)$. Suppose
the space $V^{\Lambda}(\Phi)$ is $\Gamma$-invariant and $\Phi$ are frame generators.
Then, there exists $s\geq 1$ such that the formula (\ref{eq:rank}) and the following inequality hold for a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}^d$
\begin{equation}\label{eq:frame}
s^{-1}A(\omega)\leq A(\omega)^2\leq sA(\omega).
\end{equation}
\end{theorem}
\begin{proof}
By Theorem \ref{thm:rank}, it is enough to prove that (\ref{upperlower}) implies (\ref{eq:frame}) (up to changing the constant $s$).
First observe that the previous lemma can be extended (by induction) to a sum of more than two matrices. We deduce that the lower bound in (\ref{upperlower}) implies that of (\ref{eq:frame}). The equivalence between the upper bounds follows from the fact that the number of non-zero matrices in the right-hand side of (\ref{eq:GA}) is bounded by $r$.
\end{proof}
We immediately get the following corollary.
\begin{corollary}\label{cor:frame}
Let $\Lambda$ be a closed cocompact subgroups of $\hbox{\ensuremath{\mathbb{R}}}^d$.
Assume that the space $V^{\Lambda}(\Phi)$ is translation-invariant and $\Phi$ are frame generators. Then $Tr(A(\omega))$ is not continuous, nor in $H^{\frac 12}$.
\end{corollary}
\begin{proof}
By Theorem \ref{thm:frame}, the map $\omega\to Tr(A(\omega))$ is larger than $s^{-1}$ on a set of positive (but finite) measure, and equals zero elsewhere. In particular it is not continuous and not in $H^{\frac 12}$ (by \cite{KW99}).
\end{proof}
\begin{lemma}\label{lem: constantrank}
Let $(M(x))_{x\in \hbox{\ensuremath{\mathbb{R}}}^d}$ be a continuous family of $r\times r$ non-negative self-adjoints matrices with complex coefficients. Assume that there exists $s\geq 1$ such that
$s^{-1}M(x)\leq M^2(x)$ for all $x\in \hbox{\ensuremath{\mathbb{R}}}^d$. Then the rank of $M$ is constant.
\end{lemma}
\begin{proof}
Since $x\to M(x)$ is continuous, its rank is lower semi-continuous. Let $F$ be the (closed) subset where $\mathrm{rank\ } M$ reaches its minimal value. Assume that $F$ is not open, which means that there exists a sequence $x_n\in F^c$ converging to some $x_0\in F$. It follows that the rank of $M(x_n)$ is strictly larger than the rank of $M(x_0)$, which therefore implies that the range of $M(x_n)$ intersects non-trivially the kernel of $M(x_0)$. Let $u_n$ be a sequence of unit vectors lying in this intersection. On one hand, because $u_n$ is in the range of $M(x_n)$, we must have $\|M(x_n)u_n\|\geq s^{-1}$. On the other hand, the continuity of $M$ together with the fact that the limit lies in the kernel implies $M(x_n)u_n\to 0$, contradiction. So $F$ is open which implies $F=\hbox{\ensuremath{\mathbb{R}}}^d$ hence the lemma.
\end{proof}
\begin{corollary}\label{cor:frame'}
Let $\Lambda<\Gamma$ be closed cocompact subgroups of $\hbox{\ensuremath{\mathbb{R}}}^d$.
Assume that the space $V^{\Lambda}(\Phi)$ is $\Gamma$-invariant and $\Phi$ are frame generators. Then if $A(\omega)$ is continuous, its rank is constant.
\end{corollary}
\begin{proof}The corollary follows immediately from Lemma \ref{lem: constantrank} and Theorem \ref{thm:frame}.
\end{proof}
\section{Properties of the Gramian when the $\widehat{\phi}_i$'s are continuous}\label{section:continuous}
Although this section mainly serves as preparation for Theorem \ref{epsilonthm}, we believe that it is of independent interest and could be useful elsewhere.
From now on, $\Lambda$ will always denote a lattice in $\hbox{\ensuremath{\mathbb{R}}}^d$.
The following statement is essentially trivial (and was observed for instance in \cite{Bow00, DDR94}).
\begin{proposition}\label{lem:min}
Let $\Lambda$ be a lattice in $\hbox{\ensuremath{\mathbb{R}}}^d$, and
let $\Phi=(\phi_1,\ldots, \phi_r)$ be a generating set for the $\Lambda$-invariant space $V^{\Lambda}(\Phi)$. Then the minimal number $\rho$ of generators of $V^{\Lambda}(\Phi)$ equals the essential supremum of $\mathrm{rank\ } G^{\Lambda}(\omega)$. Equivalently, if $G^{\Lambda}(\omega)$ is non invertible a.e., then one can find a generating set $\Phi'$ with $r-1$ generators.
\end{proposition}
\begin{proof}
This is an essentially trivial statement. Let $K$ be a fundamental domain for the action of $\Lambda^*$ (for example take $K=[0,1)^d$ if $\Lambda^*=\hbox{\ensuremath{\mathbb{Z}}}^d$). For every $\omega\in K$, let $V^{\Lambda}(\omega)$ be the subspace of $\ell^2(\Lambda^*)$ spanned by the $r$ vectors $v_i(\omega)=\widehat{\phi_i}(\omega+l)_{l\in \Lambda^*}$ for $i=1,\ldots,r$. Saying that $G^{\Lambda}$ is non-invertible a.e. amounts to the fact that $\dim V^{\Lambda}(\omega)\leq r-1$ for a.e. $\omega\in K$. The idea is to remove for a.e. $\omega$ one vector $v_i(\omega)$ and to recombine the other ones in order to get a new set of $r-1$ generators for $V^{\Lambda}(\Phi)$. Now, in order to do this in a measurable way, we pick the $v_i(\omega)$ of minimal index with the property that it lies in the vector space spanned by the other ones. Then we relabel the remaining ones respecting their order: for instance if $v_2$ is removed, then $v_1$ becomes $v_1'$, $v_3$ becomes $v_2'$ and so on. Now since every point of $\hbox{\ensuremath{\mathbb{R}}}^d$ can be written as $\omega+l$, for some unique $(\omega,l)\in K\times \Lambda^*$, we can define our new $r-1$ generators by the formula $\phi'_i(\omega+l)=(v'_i(\omega))_l$, for all $l\in \Lambda^*$.
\end{proof}
We deduce from this proposition that the essential supremum of $\mathrm{rank\ } G(\omega)$ equals $\rho$.
\begin{lemma}\label{lem:semicont}
Suppose that the functions $\widehat{\phi}_i$ are continuous for $i=1, \dots, r$. Then the maps
$\omega\to \mathrm{rank\ } A(\omega)$ and $\omega\to \mathrm{rank\ } G^{\Lambda}(\omega)$ from $\hbox{\ensuremath{\mathbb{R}}}^d$ to $\hbox{\ensuremath{\mathbb{N}}}$ are lower semi-continuous.
\end{lemma}
\begin{proof}
Since $G^{\Lambda}(\omega)=\sum_F A(\omega+f)$ it is enough to show that $\mathrm{rank\ } A(\omega)$ is lower semi-continuous.
Observe that $A(\omega)$ is the sum over $g\in \Gamma^*$ of the continuous positive semi-definite matrices $\widehat{\Phi}(\omega+g) \widehat{\Phi}(\omega+g)^*$. Therefore the rank of $A(\omega)$ is the supremum of the ranks of all partial finite sums. Since lower semi-continuity is stable under taking supremums, we deduce that $\mathrm{rank\ } A(\omega)$ is lower semi-continuous.
\end{proof}
\begin{proposition}\label{prop:det}
Suppose that the functions $\widehat{\phi}_i$ are continuous, and that $V^{\Lambda}(\Phi)$ is $\Gamma$-invariant for some lattice $\Gamma$ such that $[\Gamma:\Lambda]$ does not divide $\rho$. Then the subset $\{\omega, \; \mathrm{rank\ } G^{\Lambda}(\omega)<\rho\}$ is a non-empty closed subset of $\hbox{\ensuremath{\mathbb{R}}}^d$.
\end{proposition}
\begin{proof}
The fact that $\{\omega, \; \mathrm{rank\ } G^{\Lambda}(\omega)\leq \rho-1\}$ is closed results from Lemma \ref{lem:semicont}. We therefore only have to prove that it is non-empty. Let us assume on the contrary that $G^{\Lambda}(\omega)$ has rank $\rho$ for all $\omega$.
We know by Theorem \ref{thm:rank} that for a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}^d$,
\begin{equation}\label{eq:r}
\rho=\sum_{f\in F} \mathrm{rank\ } A(\omega+f).
\end{equation}
Moreover the equality is an inequality $\leq$ for all $\omega$.
But again Lemma \ref{lem:semicont} implies that the $\omega$'s for which the inequality is strict form an open set, which therefore has to be empty.
Now observe that lower semi-continuous functions with integer values can only increase locally. In other words for every $\omega_0$ the set $$\{\omega, \; \mathrm{rank\ } A(\omega)\geq \mathrm{rank\ } A(\omega_0)\}=\{\omega, \; \mathrm{rank\ } A(\omega) > \mathrm{rank\ } A(\omega_0)-1/2\}$$ is a non-empty open set. Together with (\ref{eq:r}), this immediatly implies that $\mathrm{rank\ } A(\omega)$ is locally constant, hence constant on $\hbox{\ensuremath{\mathbb{R}}}^d$ (since $\hbox{\ensuremath{\mathbb{R}}}^d$ is connected!). Let us call $m$ the corresponding integer. We therefore have $\rho=m[\Gamma:\Lambda]$, contradiction.
\end{proof}
\section{Proofs of The main results}\label{section:mainresults}
\subsection{Proof of Theorem \ref{notranslation}}
Immediately follows from Corollary \ref{cor:frame}.
\subsection{Proof of Theorem \ref{pointwise}}
Suppose by contradiction that there exists $\epsilon>0$ such that $\omega^{\frac d2+\epsilon} \widehat\phi_{i}(\omega)\in L^\infty(\hbox{\ensuremath{\mathbb{R}}}^d)$ for all $i$. This easily implies that $A(\omega)$ is continuous, so we conclude by Corollary \ref{cor:frame'}: namely since $\mathrm{rank\ } A(\omega)$ is constant it must divide the rank of $G(\omega)$, which by (\ref{eq:rank}) would imply that the index of $\Lambda$ in $\Gamma$ divides $r$.
\subsection{Proof of Theorem \ref{epsilonthm}}
The proof of Theorem \ref{epsilonthm} relies on the following result of harmonic analysis. Its proof for $d=1$ is essentially the proof of
\cite[Theorem1.2]{ASW11}. Since it extends without change to any $d$, we do not reproduce it here.
\begin{lemma}\label{epsilonlemma}
Let $\Lambda$ be a lattice in $\hbox{\ensuremath{\mathbb{R}}}^d.$
Suppose that a function $f\in L^2(\hbox{\ensuremath{\mathbb{R}}}^d)$ satisfies the following properties
\begin{enumerate}
\item there exists $\epsilon>0$ such that
$$\int_{\hbox{\ensuremath{\mathbb{R}}}^d} |f(x)|^2 |x|^{d+\epsilon} dx<\infty.$$
\item There exists a constant $C$ such that for a.e. $\omega\in \hbox{\ensuremath{\mathbb{R}}}^d$,
\begin{equation}\label{upperbond}
\sum_{k\in\Lambda^*}|\widehat f(\omega+k)|^2\le C.
\end{equation}
\item There exists $\omega_0$ is such that
$$\widehat{f}(\omega_0+k)=0 \quad {\rm for \ all}\ k\in\Lambda^*,$$
\end{enumerate}
Then for all $\eta>0$, there exists $\delta_0>0$ only depending on $\eta,\epsilon$ and $C$ such that for all $0<\delta\leq \delta_0$
$$\frac{1}{\delta^d}\int_{B(\omega_0,\delta)} \sum_{k\in\Lambda^*}|\widehat{f}(\omega+k)|^2d\omega<\eta.$$
\end{lemma}
\begin{proof}[Proof of Theorem \ref{epsilonthm}] We shall use the notation introduced in Section \ref{Section:Frame} (before stating Lemma \ref{lem:matrices}).
Let us suppose by contraction that there exits $\epsilon>0$ such that
\begin{equation}\label{epsilonequ}
\int_{\mathbb R^d} |\phi_{i}(x)|^2 |x|^{d+\epsilon} dx<\infty,
\end{equation}
for every $i$. Note that this condition implies that the $\phi_i$ are in $L^1$, so that their Fourier transforms are uniformly continuous.
Define $E=\{\omega,\;\mathrm{rank\ } G^{\Lambda}(\omega)<\rho\}$, which by Proposition \ref{prop:det}, is a non-empty closed subset of $\hbox{\ensuremath{\mathbb{R}}}^d$. Take an open ball $B\subset E^c$ whose boundary interests $E$ in a point $\omega_0$. Observe that at least one fourth of the volume of a sufficiently small ball centered in $\omega_0$ is contained in $E^c$. We will use this remark at the end of the proof.
Since $\omega_0\in E$, there exists an $(r-\rho+1)$-dimensional subspace $V_0$ of $\hbox{\ensuremath{\mathbb{R}}}^r$ on which $G^{\Lambda}(\omega_0)$ vanishes. Since $G^{\Lambda}(\omega)=F(\omega)F^*(\omega)$ where $F(\omega)$ is the $r\times |\Lambda^*|$ matrix whose $(i,l)$-coefficient is $F(\omega)_{il}=\widehat{\phi}_i(\omega+l)$, we have $F^*(\omega_0)a=0$ for all $a\in V_0$. This implies that
$$\sum_{i=1}^r a_i\widehat{\phi}_i(\omega_0+l)=0 \quad {\rm for \ all}\ l\in\Lambda^*.$$
In other words, the function $f_{a}(x)=\sum_{i=1}^ra_i\phi_i(x)$ satisfies $\widehat{f_{a}}(\omega_0+l)=0$ for all $l\in \Lambda^*$ and for all $a\in V_0$.
Now clearly the functions $f_a$ for $a\in S_{V_0}$ satisfy the assumptions of the lemma with $C$ independent of $a.$ We therefore get for $\delta>0$ small,
\begin{equation}\label{eq}
\frac{1}{\delta^d}\int_{B(\omega_0,\delta)} \sum_{k\in\widehat{\Lambda}}|\widehat{f_a}(\omega+k)|^2d\omega<\eta.
\end{equation}
Suppose $a_1, \dots, a_{r-\rho+1}$ is an orthonormal basis of $V_0$. Then for a.e. $\omega\in B$, the frame condition implies
$$s^{-1}\leq \mu^{-}(G^{\Lambda}(\omega))\leq \sum_{i=1}^{r-\rho+1}\langle G^{\Lambda}(\omega)a_i,a_i\rangle.$$
By the above remark, we see that this contradicts (\ref{eq}) provided $\eta$ is small enough.
\end{proof}
\section{Proofs of Proposition \ref{prop:regular} and Proposition \ref{prop:pointwise}}\label{section:mainpropositions}
\subsection{Proof of Proposition \ref{prop:regular}}
It is actually enough to provide a construction for $d=1$. Indeed given such $\phi_1,\ldots,\phi_{2k}$ in $L^2(\hbox{\ensuremath{\mathbb{R}}})$, we define generators in dimension $d$ by considering tensor products of these functions: namely $\tilde{\phi}_{i_1,\ldots,i_d}=\phi_{i_1}\otimes \ldots\otimes\phi_{i_d}.$ Then we get that $V_{\tilde{\Phi}}$ is simply the tensor product $V_{\Phi}^{\otimes d}$, whose generators are clearly orthonormal. Translation invariance of $V_{\tilde{\Phi}}$ therefore follows from that of $V_{\Phi}$.
Hence, let us focus on the one-dimensional case.
Let $g$ be an infinitely-differentiable function that satisfies $g(x)=0$ when $x\le 0$, $g(x)=1$ when $x\ge 1$,
and $g^2(x)+g^2(1-x)=1$ when $0\le x\le 1$. Define $\widehat{\phi}_1(\omega)=g(\omega)g(2-\omega)$ and $\widehat{\phi}_2(\omega)=g(1-\omega)\chi_{[0,1)}-g(\omega-1)\chi_{[1,2)}$.
Then for $i\le r$, define $\widehat{\phi}_i(\omega)=\widehat{\phi}_1(\omega-i+1)$ when $i$ is odd and $\widehat{\phi}_i(\omega)=\widehat{\phi}_2(\omega-i+2)$ when $i$ is even ($\widehat{\phi}_1$ and $\widehat{\phi}_2$ are plotted in Figure 1). It is easy to check that $\sum_{l\in \hbox{\ensuremath{\mathbb{Z}}}}|\widehat{\phi}_i(\omega+l)|^2=1$ a.e. $\omega$ for all $i$, and that $\sum_{l\in \hbox{\ensuremath{\mathbb{Z}}}}\widehat{\phi}_1(\omega+l)\overline{\widehat{\phi}_2}(\omega+l)=0$. We deduce that $\sum_{l\in \hbox{\ensuremath{\mathbb{Z}}}}\widehat{\phi}_i(\omega+l)\overline{\widehat{\phi}_j}(\omega+l)=0$ for $i\neq j$. This shows that $G(\omega)= I$ which amounts to saying that the functions $\phi_i$ are orthonormal generators for $V(\Phi)$.
\begin{figure}[hbt]
\centering
\begin{tabular}{c}
\includegraphics[width=80mm]{phii2.eps}
\end{tabular}
\caption{The function $\widehat{\phi}_1$ being red and the function $\widehat{\phi}_2$ being blue.
}
\label{figure1}
\end{figure}
Since all the functions are supported on $[0,r]$, we can show that, for a.e. $\omega\in[0,1)$, $\mathrm{rank\ }A(\omega+f)=1$ when $0\le f\le r-1$ and $\mathrm{rank\ }A(\omega+f)=0$ elsewhere. Hence $\sum_{f\in \hbox{\ensuremath{\mathbb{Z}}}}\mathrm{rank\ }A(\omega+f)=r=\mathrm{rank\ } G^{\Lambda}(\omega)$. From Corollary \ref{cor:rank}, $V(\Phi)$ is translation-invariant.
Observe that when $i$ is odd, $\widehat{\phi}_i$ is compactly supported and infinitely-differentiable. When $i$ is even, $\widehat{\phi}_i$ can be written as a product of a compactly supported and infinitely-differentiable function and the characteristic function $\chi_{[i-2,i)}$. Since $\chi_{[i-2,i)}$ belongs to $H^{\frac12-\epsilon}$, we have $\widehat{\phi}_i\in H^{\frac12-\epsilon}$ for any $\epsilon>0$.
\subsection{Proof of Proposition \ref{prop:pointwise}}
We assume that $\Gamma=\frac{1}{n_1}\hbox{\ensuremath{\mathbb{Z}}}\times\frac{1}{n_2}\hbox{\ensuremath{\mathbb{Z}}}\times\cdots\times\frac{1}{n_d}\hbox{\ensuremath{\mathbb{Z}}}$.
We first need the following lemma which treats the case when $d=1$ and $r=1$.
\begin{lemma}[\cite{ASW11}]\label{dimension1}
For integer $n_1$, there exists a function $\phi_1 \in L^1\cap L^2$ (and hence $\widehat \phi_1$ is continuous), such that $\phi_1$ is an orthonormal generator for its generating space $V(\phi_1)$,
$V(\phi_1)$ is $\frac{1}{n_1}\hbox{\ensuremath{\mathbb{Z}}}$-invariant and $\omega^{\frac 12} \widehat\phi_1(\omega)\in L^\infty(\hbox{\ensuremath{\mathbb{R}}})$.
\end{lemma}
Since our method relies on the construction of $\phi_1$ in Lemma \ref{dimension1}, we will describe $\phi_1$ explicitely here. Let $g$ be the function as in the proof of Proposition \ref{prop:regular}. Define $ g_0(x)= g(x+1) g(-x+1)$ and $g_1(x)=g(x+1)g(-2x+1)$. Then the Fourier transform of $\phi_{1_1}$ is defined to be
\begin{eqnarray}\label{time.lem.eq1}
\widehat {\phi_1}(\omega) & = & h_0(\omega)+\sum_{j=1}^\infty \sum_{l=0}^{4^j-1} 2^{-j} h_j(\omega-n_1 (\gamma_j+l))\nonumber\\
& & \quad +\sum_{j=1}^{\infty} \sum_{l=0}^{4^j-1} 2^{-j}
h_j(-\omega-n_1 (\gamma_{j}+l)),
\end{eqnarray}
where $\gamma_j=\sum_{k=0}^{j-1} 4^k$, $h_0(\omega)=g_0(4\omega)$ and $h_j(\omega)=g_1(2^{j+1}\omega-2^j+1)$.
It is not hard to check that $\widehat {\phi_1}$ is an even function such that $|\widehat\phi_1|\le1$ and has support on $[-\frac14,\frac14]\cup( \cup_{j=1}^\infty E_{1j}\cup E_{1j}^\prime)$, where $E_{1j}=\cup_{l=0}^{4^j-1}[\frac12-\frac1{2^j}+n_1(\gamma_j+l), \frac12-\frac1{2^{j+2}}+n_1(\gamma_j+l)]$ and $E_{1j}^\prime=\cup_{l=0}^{4^j-1}[\frac1{2^j}-\frac12-n_1(\gamma_j+l), \frac1{2^{j+2}}-\frac12-n_1(\gamma_j+l)]$.
\noindent{\bf From one to several generators (in dimension 1).} Now we want to construct a $\frac{1}{n_1}\hbox{\ensuremath{\mathbb{Z}}}$-invariant SIS with $r$ orthonormal generators in $L^1\cap L^2$ satisfying our pointwise decay property in Fourier domain.
Define $\phi_{i}$ for $2\le i\le r$ to be such that $\widehat {\phi}_{i}(\omega)=0$ when $\omega\in[0,n_1\gamma_{2i-2}-1]$ and
\begin{align}
\widehat {\phi}_{i}(\omega)= \frac{1}{\sqrt2}\Big(\sum_{l=0}^{4^{2i-2}-1}2^{-(2i-2)}h_0(\omega-n_1(\gamma_{2i-2}+l))\Big)\nonumber\\
+\sum_{j=1}^\infty \sum_{l=0}^{4^{j+2i-2}-1} 2^{-(j+2i-2)} h_j(\omega-n_1(\gamma_{j+2i-2}+l))
\end{align}
when $\omega\ge n_1\gamma_{2i-2}-1$, and $\widehat \phi_{i}(\omega)= \widehat \phi_{i}(-\omega)$ when $\omega\le0$.
It is easy to see that each $\widehat\phi_{i}$ obeys the pointwise decay property and that it is an orthonormal generator for the principal SIS $V(\phi_{i})$. In fact,
$$\sum_{l\in {\mathbb Z}}|\widehat \phi_{i}(\omega+l)|^2=\sum_{l\in {\mathbb Z}}|\widehat \phi(\omega+l)|^2=1$$
for a.e. $\omega$. One can also check that for $2\le i\le r$, $\widehat\phi_{i}$ has support
\begin{align}
\Big( \cup_{j=1}^\infty E_{ij}\cup E_{ij}^\prime\Big)\bigcup\Big(\cup_{l=0}^{4^{2i-2}-1}[-\frac14-n_1(\gamma_{2i-2}+l),\frac14-n_1(\gamma_{2i-2}+l)]\Big)\nonumber\\
\bigcup\Big(\cup_{l=0}^{4^{2i-2}-1}[-\frac14+n_1(\gamma_{2i-2}+l),\frac14+n_1(\gamma_{2i-2}+l)]\Big),
\end{align}
where $E_{ij}=\cup_{l=0}^{4^{j+2i-2}-1}[\frac12-\frac1{2^j}+n_1(\gamma_{j+2i-2}+l), \frac12-\frac1{2^{j+2}}+n_1(\gamma_{j+2i-2}+l)]$ and $E_{ij}^\prime=\cup_{l=0}^{4^{j+2i-2}-1}[\frac1{2^j}-\frac12-n_1(\gamma_{j+2i-2}+l), \frac1{2^{j+2}}-\frac12-n_1(\gamma_{j+2i-2}+l)]$.
In particular, each $V(\phi_{i})$ is $\frac1{n_1}$-invariant and all the $\phi_{i}$'s have disjoint support in Fourier domain. Hence $V(\Phi_1)$ is also $\frac1{n_1}$-invariant, where $\Phi_1$ is the column vector whose components are the $\phi_{i}$'s. In fact
\begin{equation}\label{orthognal}
V(\Phi_1)=\bigoplus_{i\le r}V(\phi_{i})
\end{equation}
From Lemma \ref{dimension1}, for each $j\ge2$, we can construct $V(\psi_j)$ with orthonormal generator $\psi_j$ having the desired properties and that $V(\psi_j)$ is $\frac1{n_j}\hbox{\ensuremath{\mathbb{Z}}}$-invariant.
\noindent{\bf Higher dimension.}
Like in the previous section, we let $\tilde{\phi}_{i}$ be the $d$-fold tensor product $\phi_{i}\otimes\psi_2 \ldots\otimes\psi_d$, for $i=1,\ldots, r$. By construction, the shift-invariant space generated by these $r$ orthonormal function is $\Gamma$-invariant. Since all $\phi_{i}$'s and $\psi_{j}$'s are in $L^1(\hbox{\ensuremath{\mathbb{R}}})$, it follows that the $\tilde{\phi}_{i}$'s are in $L^1(\hbox{\ensuremath{\mathbb{R}}}^d)$. Pointwise decay results from the fact that all these functions have bounded Fourier transforms.
\bigskip
{\bf Acknowledgement} \quad The authors would like to thank Akram Aldroubi and Qiyu Sun for valuable discussions. The authors would also like to thank Carolina Mosquera and Victoria Paternostro for their comments on the proof of Theorem 1.2.
|
\section{Introduction}
The Maxwell algebra as the enlargement of Poincar\'{e} algebra ($P_\mu, M_{\mu\nu}$) by antisymmetric Abelian tensorial charges $Z_{\mu\nu} =-Z_{\nu\mu}$ was firstly obtained for $D=4$ more than forty years ago \cite{Bacry:1970ye,Schrader:1972zd} by supplementing the relations
\begin{eqnarray}\label{wroja1.1}
[P_\mu, P_\nu ] &= & i\, M^2 \, Z_{\mu\nu}\, ,
\label{wroja1a}
\nn\\
\left[ P_{\rho}, Z_{\mu\nu} \right] & = & 0, \qquad \qquad
[ M_{\rho\tau} , Z_{\mu\nu} ] =-i(\h_{\mu[\tau}\,Z_{\rho]\nu}-
\h_{\nu[\tau}\,Z_{\rho]\mu})\,,
\label{wroja1b}
\end{eqnarray}
where $M$ is a geometric mass parameter ($[M]=1;$ we use in $D=4$ the metric $\eta_{\mu\nu}={\rm diag}(-1,1,1,1))$.
In presenting the Maxwell algebra we introduce mass-like fundamental parameter $M$ (in \cite{biwroja3} denoted by $\Lambda=M^2$) which implies vanishing mass dimensionality of the tensorial charges $([Z_{\mu\nu}]=0)$. The relations (\ref{wroja1a}) can be however rewritten for generators $Z_{\mu\nu}$ with any mass dimensionality. In particular if $M^2$ is replaced by electromagnetic dimensionless coupling constant $e$ ($[e]=0$; see e.g. \cite{Bacry:1970ye,Schrader:1972zd,biwroja4new,Gomis:2009vm}) we obtain $[Z_{\mu\nu}]=2$. Simple dynamical realization of such Maxwell algebra is obtained in the model of relativistic free particle moving in constant EM field backgrounds \cite{biwroja4new,Gomis:2009vm}
We recall that Maxwell algebra can be obtained by a contraction of $O(3,1)\oplus O(3,2)$ \cite{Soroka:2006aj,Gomis:2009vm}\footnote{Recently deformation $O(3,1)\oplus O(3,2)$ of Maxwell algebra was called $AdS$-Maxwell algebra \protect\cite{biwroja7}.} or $O(3,1)\oplus O(4,1)$ \cite{Gomis:2009vm} (if we replace $M^2 \to - M^2$ in \bref{wroja1.1}).
Recently in \cite{Bonanos:2009wy} there were as well introduced the simple Maxwell superalgebras, with two Weyl supercharges ${\bf Q}_\alpha , {\bf \Sigma}_\alpha$, where ${\bf Q}_\alpha$ are the standard $N={1}$ Poincar\'{e} supercharges and new Maxwell supercharges ${\bf \Sigma}_\alpha$ are required for the supersymmetrization of the generators $Z_{\mu\nu}$. One should point out that recently appeared proposals to use the Maxwell algebra \cite{biwroja3,biwroja9,biwroja7,Durka:2011va} and Maxwell superalgebra \cite{biwroja10} to the geometric extension of (super)gravity theories.
At present it appears interesting to study the problem how the Maxwell algebra can be supersymmetrized in various ways. In order to obtain important class of $D=4$ Maxwell superalgebras we observe that $D=4$ Lorentz algebra can be obtained as the realification\footnote{
By realification of n-dimensional complex Lie (super)algebra $\widehat{g}$ with the generators $g_i=g^{(1)}_i + i g^{(2)}_i, (i=1,...,n)$ we call the 2n-dimensional real Lie (super)algebra $\widehat{g}_R$ {on real space} with real generators $g^{(1)}_i, g^{(2)}_i$ (see e.g. \cite{Onishchik:1991}).
If besides the complex generators $g_i$ we introduce $\ba g_i=g^{(1)}_i - i g^{(2)}_i$, the real generators of $\8g_R$ are linear combinations of
$g_i$ and $\ba g_i$, ($g^{(1)}_i=\frac12(g_i+\8g_i),\;g^{(2)}_i=\frac1{2i}(g_i-\8g_i)).$}
of complex algebra $Sp(2|C)$ ($O(3,1)\simeq Sp_R(2|C)$) with the supersymmetrization described by the realification of complex superalgebra $OSp(m;2|C)$,
(for $m=1$ see e.g. \cite{biwroja11}); further it is well known that $D=4$ $AdS$ symmetries are supersymmetrized by $D=4$ $AdS$ superalgebras $OSp(k;4)$.
We shall consider therefore contractions\footnote{
We need to consider the contractions of direct sum $\hat{g}_1 \oplus \hat{g}_2$ of Lie (super)algebras in which the contracted generators appear as suitable sums of the generators belonging to $\hat{g}_1$ and $\hat{g}_2$.
For earlier discussion of contractions which use the linear superposition of generators from the sum of Lie (super)algebras, see \cite{biwroja12,biwroja13,biwroja14}.
}
of the following choices of the sum of real semisimple superalgebras
\begin{equation}
\label{wroja1.3} O(3,1)\oplus O(3,2)\quad \longrightarrow^{\hskip -9mm {SUSY}}\quad OSp_R(2N-k; 2|C) \oplus OSp(k;4)
\end{equation}
where $k=0,1,2\ldots 2N$ and $N=0,1,2\ldots$.
We assume that the derived extended Maxwell superalgebras should satisfy the following two properties:
\begin{description}
\item{i)} They should contain at least one standard bilinear fermionic SUSY relation characterizing Poincar\'e supercharges with the following algebraic structure
\begin{equation}
\label{wroja1.4}
\left\{ \bQ_{\alpha}, {\bQ}_{{\beta}} \right\}
= (C\gamma^{\mu} \, P_{\mu})_{\alpha {\beta}}+{\rm\;central\; charges}.
\end{equation}
\item{ii)} Their bosonic sector should contain the Maxwell algebra.
\end{description}
Let us observe that in the contraction of \bref{wroja1.3}
the Poincar\'{e} supercharges satisfying the relations \bref{wroja1.4}
can be only obtained from the contraction of $OSp(k;4)$.
If we put $k=0$ in \bref{wroja1.3}
the generators $Sp(4)\simeq O(3,2)$ remain not supersymmetrized and we can get by contraction only a purely exotic version of Maxwell superalgebra \cite{biwroja15,biwroja13} with the standard generic SUSY relations \bref{wroja1.4} replaced by the following one
\begin{equation}
\label{wroja1.5}
\{ {\bf S}_{\alpha}, {\bf S}_{\beta} \}
= (C\gamma^{\mu\nu} Z_{\mu\nu})_{\alpha\beta}\,+{\rm\;central\; charges}.
\end{equation}
The paper is organized as follows: In Sect.~2 we shall derive the simple $N=1$ Maxwell superalgebras. If we put $N=1$ in \bref{wroja1.3}, for $k=2$ and $k=1$ we shall obtain by contractions two versions of simple ($N=1$) $D=4$ Maxwell superalgebras, differing by the presence of Abelian chiral generator (see also \cite{Bonanos:2009wy,biwroja14}).
In Sect.~3 we consider the nonstandard contraction of \bref{wroja1.3} in general case (arbitrary $N$ with $k=1,2,\ldots,2N$); we treat
separately the cases $0\leq k< N$ and $N\leq k\leq 2N$ requiring different explicit contractions. In Sect~4 we provide short discussion of our results.
\section{$N=1$ Maxwell superalgebras by contraction}
\subsection{Simple Maxwell superalgebras}
In this section we consider the contractions of the superalgebras \bref{wroja1.3}
for $N=1$ and $k=1,2$ providing the simple Maxwell superalgebras with two Weyl supercharges which were proposed in
\cite{Bonanos:2009wy}. The Maxwell algebra (relations (\ref{wroja1a})+Lorentz algebra) is extended supersymmetrically as follows\footnote{
The mass dimension of the generators are $[P_{\mu}]=1,\;[Z_{\mu\nu}]=0,\; [\bQ_\A]=\frac12,\; [\bS_\A]=\frac32,\; [B]=2,\; [B_5]=0$.}
\bea
\{{\bQ}_\A,{\bQ}_\B\}&=&(C\gamma^{\mu})_{\A\B}P_\mu,\qquad
\{{\bSig}_\A,{\bSig}_\B\}=0,
\nn\\
\{{\bQ}_\A,{\bf\Sigma}_{\B}\}&=&\frac{M^2}2(C\gamma^{\mu\nu})_{\A\B}\,Z_{\mu\nu}\,+\, (C\gamma_5)_{\A\B}\,{\rB},
\label{Maxsuperalgebra-1}
\\
\left[P_\mu,{\bQ}_\A\right]&=&-\frac{i}2\,\bSig_{\B}(\gamma_{\mu}{)^{\B}}_\A,\qquad
\left[P_\mu,{\bSig}_\A\right]=0,
\nn\\
\left[M_{\rho\s},\bQ_{\A}\right]&=&-\frac{i}2(\bQ \gamma_{\rho\s})_{\A},\qquad
\left[M_{\rho\s},\bSig_{\A}\right]=-\frac{i}2(\bSig \gamma_{\rho\s})_{\A}\,,
\label{Maxsuperalgebra-2}
\\
\left[B_C,{\bQ}_\A\right]&=&i\,(\bQ\gamma_{5})_\A,\qquad
\left[B_C,{\bSig}_\A\right]=-\,i\,(\bSig\gamma_{5})_\A\,,
\label{Maxsuperalgebra}
\eea
where real Dirac-Majorana matrices $\gamma_{\mu}$
($\gamma^T_i=\gamma_i, \gamma^T_0=-\gamma_0$) verify
the $O(3,1)$ Clifford algebra $\{ \gamma_\mu, \gamma_\nu \}=2\eta_{\mu\nu}$ and $\gamma_{\mu\nu}=\half [ \gamma_\mu, \gamma_\nu ]= -\gamma_{\nu\mu}$ is the $4\times 4$ matrix realization of $O(3,1)$.
The charge conjugation matrix $C=\gamma_0$ and $\gamma_5 = \gamma_0 \gamma_1 \gamma_2 \gamma_3 = - \gamma^T_5$ satisfy the relations
$(C\gamma_\mu)^T=(C\gamma_\mu),\;(C\gamma_{\mu\nu})^T=(C\gamma_{\mu\nu}),\;
(C\gamma_5)^T=-(C\gamma_5).
$ The new supercharges ${\bf \Sigma}_\alpha$ are needed for the supersymmetrization of the generators $Z_{\mu\nu};$ the $B_C$ describes a chiral generator and $B$ is the central charge. In the minimal $N=1$ Maxwell superalgebra we put $B=B_C=0$.
\subsection{Contraction of $O(3,1) \oplus OSp(2;4)\;\;(k=2$ case)}
The $OSp(2,4)$ superalgebra is
\bea
\left[\CM_{{\8\mu}{\8\nu}},\CM_{\8\rho\8\s}\right]&=&
-i\,\h_{\8\rho[{\8\nu}}\CM_{{\8\mu}]\8\s}+i\,
\h_{{\8\s}[\8\nu}\CM_{{\8\mu}]\8\rho},
\label{OSp04}\\
\{{\CQ}_{\A}^{ i},{\CQ}_{\B}^{ j}\}&=&-\,\D^{ij}\,(C\Gam^{{\8\mu}{\8\nu}})_{\A\B}
\CM_{{\8\mu}{\8\nu}}+\,2\,(C)_{\A\B}\CB^{ij},
\nn\\
\left[\CM_{{\8\mu}{\8\nu}},\CQ_{\A}^{ i}\right]&=&-\frac{i}2(\CQ^i \Gam_{{\8\mu}{\8\nu}})_{\A},
\qquad
\left[\CB^{ij},{\CQ}_{\A}^{k}\right]=-i\,\D^{k[j}\,\CQ_{\A}^{ i]}.
\label{OSp24}\eea
Here $ \CM_{{\8\mu}{\8\nu}},\, (\8\mu=0,1,2,3,4) $ are $SO(3,2)$ generators with $\h_{{\8\mu}{\8\nu}}=(-1,1,1,1,-1)$ and
$O(3,1)$ real Dirac-Majorana matrices
$\Gamma_\mu = \gamma_{\mu}\gamma_{5}$, $\Gamma_4= -\gamma_5$
satisfy the $O(3,2)$ Clifford algebra
$\{\Gamma_{\hat{\mu}}, \Gamma_{\hat{\nu}} \}= 2\eta_{\hat{\mu} \hat{\nu}}$.
The $4\times 4$ matrix realization $\Gamma_{\hat{\mu}\hat{\nu}}=\half[ \Gamma_{\hat{\mu}}, \Gamma_{\hat{\nu}} ]$ of $O(3,2)$ algebra is expressed by
$O(3,1)$ $\gamma_\mu$-matrices $ as
\Gamma_{\mu\nu} = \half [ \Gamma_\mu , \Gamma_\nu ]
= \gamma_{\mu\nu}, \;
\Gamma_{\mu4}= \Gamma_\mu \Gamma_4 = \gamma_\mu.$
The charge conjugation $C=\gamma_0 = \Gamma_0 \Gamma_4$ is common
for $O(3,2)$ and $O(3,1)$.
Supercharges ${\CQ}_{\A}^{i},\,(\A=1,2,3,4,\,i=1,2)$ are real $O(3,2)$ spinors and $\CB^{ij}=-\ep^{ij}\CB_C$ is the $SO(2)$ generator.
The real algebra $O(3,1)$ is provided by
the algebra $Sp_R(2|C)=O_R(2,1|C) \simeq O(3,1)$.
The algebra $O(2,1|C)\oplus{\overline{O(2,1|C)}}$ is described by the generators $\bJ_{\ba\mu}$ and $\bJ^\dagger_{\ba\mu}, ({\ba\mu}=0,1,2)$ with the metric $\h_{{\bar\mu}{\bar\nu}}=(-1,1,1)$,
\bea
\left[\bJ_{\ba\mu},\bJ_{\ba\nu}\right]={i}\,{\ep_{\ba\mu\ba\nu}}^{\ba\rho}
\bJ_{\ba\rho},\qquad \left[\bJ^\dagger_{\ba\mu},\bJ^\dagger_{\ba\nu}\right]={i}\,{\ep_{\ba\mu\ba\nu}}^{\ba\rho}\bJ^\dagger_{\ba\rho},\qquad
[\bJ_{\ba\mu},\bJ^\dagger_{\ba\nu}]=0,
\label{OSp02}\eea
where $\ep^{\ba\mu\ba\nu\ba\rho}$ is the Levi-Civita symbol with $\ep^{012}=-\ep_{012}=1$.
If we introduce the real $O(3,1)$ generators
$\CJ_{\mu\nu}, \,(\mu=0,1,2,3)$ by the relation
\be
\bJ_{\ba\mu}=\frac14{\ep_{\ba\mu}}^{\ba\rho\ba\nu}
J_{\ba\rho\ba\nu}+\frac{i}2\,J_{\ba\mu 3}\,,
\label{bJdef}
\ee
they satisfy the $D=4$ Lorentz algebra with $\h_{\mu\nu}=(-1,1,1,1)$,
\bea
\left[\CJ_{{\mu}{\nu}},\CJ_{\rho\s}\right]&=&-i\,
\h_{{\rho}[\nu}\CJ_{{\mu]}\s}+i\,\h_{{\s}[\nu}\CJ_{{\mu]}\rho}\,.
\label{O31}\eea
We propose the following redefinitions of the generators of $O(3,1)\oplus O(3,2)$ (we recall that $\Gamma_4=-\gamma_5$ and
we put $\A+\B=1$) \be
M_{\mu\nu}= \CJ_{\mu\nu} + \CM_{\mu\nu},\quad
P_\mu = \frac1R\,{\CM_{\mu 4}},\quad
Z_{\mu\nu} = \frac{1}{R^2 M^2} (\A\CJ_{\mu\nu}-\B{\CM}_{\mu\nu})
\label{relgenSM3Rplus-1}
\ee
and the rescaled internal symmetry generator and supercharges as follows
\bea
B_C&=&\frac1{R^{\gamma}} {\CB}_C\,, \label{relgenSM3RplusB}
\\
{\bf Q}_{\alpha} = \frac{1}{2R^{1/2}} (Q^1_\alpha+( Q^2\gamma_5 )_{\alpha}),
&\,& d
{\bf \Sigma}_{\alpha} = \frac{1}{2R^{3/2}} (Q^1_\alpha-(Q^2\gamma_5)_{\alpha}) \,,\label{relgenSM3Rplus}
\eea
where $R$ in the formulas \bref{relgenSM3Rplus-1}-\bref{relgenSM3Rplus}
is a contraction parameter with dimension of length $([R]={-1})$. The superalgebra $O(3,1)\oplus OSp(2;4)$ in terms of new rescaled real generators takes the form
\bea
\left[P_\mu,P_{\nu}\right]&=&i\,M^2\,Z_{\mu\nu}-i\frac{\B}{R^2}\,M_{\mu\nu},\qquad
\left[P_\mu,M_{\rho\s}\right]=-i\,\h_{\mu[\rho}P_{\s]},\nn\\
\left[P_\mu,Z_{\rho\s}\right]&=&i\,\frac{\A}{R^2}\,\h_{\mu[\rho}P_{\s]},\qquad
\left[M_{\mu\nu},Z_{\rho\s}\right]=-i\,\h_{\rho[\nu}Z_{\mu] \s}+i\,
\h_{\s[\nu}Z_{\mu]\rho},\nn\\
\left[M_{\mu\nu},M_{\rho\s}\right]&=&-i\,\h_{\rho[\nu}M_{\mu] \s}+i\,
\h_{\s[\nu}M_{\mu]\rho},\nn\\
\left[Z_{\mu\nu},Z_{\rho\s}\right]&=&-i\frac{\B-\A}{M^2\,R^2}
(\,\h_{\rho[\nu}Z_{\mu] \s}-\,\h_{\s[\nu}Z_{\mu]\rho})
-i\frac{\B\A}{M^4\,R^4}(\h_{\rho[\nu}M_{\mu] \s}-\,
\h_{\s[\nu}M_{\mu]\rho}),\label{BMaxalgeb02}
\eea\bea
\{{\bQ}_\A,{\bQ}_\B\}&=&(C\gamma^{\mu})_{\A\B}P_\mu,\qquad
\{{\bSig}_\A,{\bSig}_\B\}=\frac1{R^2}\,(C\gamma^{\mu})_{\A\B}P_\mu,
\nn\\
\{{\bQ}_\A,{\bf\Sigma}_{\B}\}&=&\frac{M^2}2(C\gamma^{\mu\nu})_{\A\B}\,Z_{\mu\nu}\,+\,\frac1{R^{2-\gam}}\, (C\gamma_5)_{\A\B}\,B_C,
\\
\left[P_\mu,{\bQ}_\A\right]&=&-\frac{i}2\,\bSig_{\B}(\gamma_{\mu}{)^{\B}}_\A,\qquad
\left[P_\mu,{\bSig}_\A\right]=-\frac{i}{4R^2}\,\bQ_{\B}(\gamma_{\mu}{)^{\B}}_\A,
\nn\\
\left[Z_{\mu\nu},{\bQ}_\A\right]&=&\frac{i}{2R^2 M^2}\,(\bQ\gamma_{\mu\nu})_\A,\qquad
\left[Z_{\mu\nu},{\bSig}_\A\right]=\frac{i}{2R^2 M^2}\,(\bSig\G_{\mu\nu})_\A,
\nn\\
\left[M_{\rho\s},\bQ_{\A}\right]&=&-\frac{i}2(\bQ \gamma_{\rho\s})_{\A},\qquad
\left[M_{\rho\s},\bSig_{\A}\right]=-\frac{i}2(\bSig \gamma_{\rho\s})_{\A}.
\nn\\
\left[B_C,{\bQ}_\A\right]&=&\frac{i}{R^{\gam}}\,(\bQ\gamma_{5})_\A,\qquad
\left[B_C,{\bSig}_\A\right]=-
\, \frac{i}{R^{\gam }}\,(\bSig\gamma_{5})_\A \,.
\label{kalgebra}
\eea
If we put $\A=1, \B=0, $ \,
$2>\gam >0$ and $M=1$ the formulas \bref{BMaxalgeb02}-\bref{kalgebra} describe the k-deformation
of the Maxwell superalgebra
given in \cite{biwroja14}($\frac{1}{R^2}=$k), which reproduces in the k$\to 0$ limit the Maxwell superalgebra (\ref{Maxsuperalgebra-1}-\ref{Maxsuperalgebra}) with $B_C=0$ and $B=0$.
In the case $\gamma=0$, $B_C$ is nonvanishing and becomes a chiral charge generating
chiral transformation of $\bQ_\A$ and $\bSig_\A$, (see \bref{Maxsuperalgebra}).
The specific feature of our contraction is that two factors from the defining algebra $OSp(2|4)\oplus \SO(3,1)$ are suitably mixed (see \bref{relgenSM3Rplus-1}--\bref{relgenSM3Rplus}).
As a result, this type of contractions will not respect the direct sum structure of the uncontracted algebras
(see also \cite{Cangemi:1992ri},\cite{deAzcarraga:2002xi} (Sect.~8)).
\subsection{Contraction of $OSp_R(1; 2|C) \oplus OSp(1;4)\;\; (k=1$ case) }
The $OSp(1;4)$ superalgebra is
\bea
\left[\CM_{{\8\mu}{\8\nu}},\CM_{\8\rho\8\s}\right]&=&
-i\,\h_{{\8\rho}[\8\nu}\CM_{{\8\mu}]\8\s}+i\,
\h_{{\8\s}[\8\nu}\CM_{{\8\mu}]\8\rho},
\nn\\
\{{\CQ}_{\A}^{},{\CQ}_{\B}^{}\}&=&-\frac{1}{2}\,(C\Gam^{{\8\mu}{\8\nu}})_{\A\B}
\CM_{{\8\mu}{\8\nu}},\qquad
\left[\CM_{{\8\mu}{\8\nu}},\CQ_{\A}\right]=-\frac{i}2(\CQ\Gam_{{\8\mu}{\8\nu}})_{\A}.
\eea
The complex superalgebra $OSp(1;2|C)=(\bJ_{\ba\mu},{ S}_{+\A})$ contains the bosonic subalgebra
$Sp(2|C)=O(2,1|C)$ with its realification
given by $O(3,1)$. The complex supercharges ${ S}_{+\A}$ define the
fermionic sector of $OSp(1;2|C)$,
\bea
\{ { S}_{+\A},{ S}_{+\B}\}
&=&4i\,(C\G_+^{\ba\rho})_{\A\B}\,\bJ_{\ba\rho},
\qquad
\left[\bJ_{\ba\mu},{ S}_{+\A}\right]=\frac12\,{{
S}_{+\B}\,({\Gamma_{+\ba\mu}})^\B}_\A,
\label{OSp12C}\eea
where the matrices ${\Gamma_{+\ba\mu}}$ are defined by O(3,1) gamma matrices $\gamma_\mu$ as
\bea
{\G}_{+\ba\mu}&=&\gamma_{\ba\mu}\gamma_3\,P_+,\quad P_\pm=\frac12(1\pm i\,\gamma_5),\quad
[{\G}_{+\ba\mu},{\G}_{+\ba\nu}]=2i\,{\ep_{\ba\mu\ba\nu}}^{\ba\rho}{\G}_{+\ba\rho}
\eea and ${ S}_{+\A}={ S}_{+\A}P_+$.
The complex conjugated superalgebra
${\overline{OSp(1;2|C)}}=({ S}_{-\B}={S}^\dagger_{+\B},\bJ_{\ba\mu}^\dagger )$
has the following fermionic sector
\bea
\{{S }_{-\A},{ S}_{-\B}\}&=&-4i\,(C\G_-^{\ba\rho})_{\A\B}\,\bJ^\dagger_{\ba\rho},\qquad
\left[\bJ^\dagger_{\ba\mu},{ S}_{-\A}\right]=-\frac12\,
{{\bf S}_{-\B}\,({\Gamma_{-\ba\mu}})^\B}_\A.
\label{OSp12Cm}\eea
Using \bref{bJdef} and introducing real supercharges
${ \CS}_\A={S}_{+\A}+{ S}_{-\A}$
one gets the following real superalgebra describing the realification $OSp_R(1;2|C)$ of $OSp(1;2|C)$ which extends supersymmetrically the Lorentz algebra \bref{O31} by the relation
\be
\{{{\CS }}_{\A},{{\CS}}_{\B}\}=-\,(C\gamma^{{\mu}{\nu}})_{\A\B}\CJ_{{\mu}{\nu}},
\qquad
\left[\CJ_{{\mu}{\nu}},{\CS}_{\A}\right]=-\frac{i}2(\CS \gamma_{{\mu}{\nu}})_{\A}.
\label{OSp14alg22}\ee
The superalgebra $OSp_R(1; 2|C) \oplus OSp(1;4)$ before the contraction $R\to \infty$ is described by the real generators $ (\CM_{\8\mu\8\nu},\CQ_\A,\CJ_{\mu\nu},\CS_\A)$ and does not contain any internal symmetry generators. Using
the rescalings \bref{relgenSM3Rplus-1} of $O(3,1)\oplus O(3,2)$ generators and new rescaled supercharges
\be
{\bQ}_{\alpha}={R^{-1/2}}({\CQ}_{\alpha} +{\CS}_{\alpha}),\quad
{\bSig}_{\alpha}={R^{-3/2}}{\CS}_{\alpha},
\label{comgeb}
\ee
we describe the real superalgebra $OSp_R(1; 2|C)\oplus OSp(1|4)$. The bosonic part is given by the relations
\bref{BMaxalgeb02} and the part of superalgebra which contains the supercharges $\bQ_\A$ and $\bSig_\A$ is
\bea
\left[P_\mu,{\bQ}_\A\right]&=&\frac{i}2\,(\bSig\gamma_{\mu})_\A
-\frac{i}{2R}\,(\bQ\gamma_{\mu})_\A,\qquad
\left[P_\mu,{\bSig}_\A\right]=0,
\label{kalgebraPQ2}
\nn\\
\left[Z_{\mu\nu},{\bQ}_\A\right]&=&-\frac{i}2\frac1{R M^2}\,(\bSig\gamma_{\mu\nu})_\A,\qquad
\left[Z_{\mu\nu},{\bSig}_\A\right]= -\frac{i}2\,\frac1{R^{2}M^2}\,(\bSig\gamma_{\mu\nu})_\A,
\nn\\
\left[M_{\rho\s},\bQ_{\A}\right]&=&-\frac{i}2(\bQ \gamma_{\rho\s})_{\A},\qquad
\left[M_{\rho\s},\bSig_{\A}\right]=-\frac{i}2(\bSig \gamma_{\rho\s})_{\A}\,,
\\
\{{\bQ}_\A,{\bQ}_\B\}&=&-\frac{1}R\,(C\gamma^{\mu\nu})_{\A\B}\,M_{\mu\nu}\,+
2\,(C\gamma^{\mu})_{\A\B}P_\mu,
\nn\\
\{{\bQ}_\A,{\bf\Sigma}_{\B}\}&=&-{M^2}\,(C\gamma^{\mu\nu})_{\A\B}Z_{\mu\nu},
\qquad
\{{\bSig}_\A,{\bSig}_\B\}=-\,\frac{M^2}R\,(C\gamma^{\mu\nu})_{\A\B}Z_{\mu\nu}.
\label{kalgebra2}
\eea
In the limit $R\to\infty$ the superalgebra
$OSp_R(1; 2|C) \oplus OSp(1;4)$ is contracted into the minimal simple Maxwell superalgebra (see \cite{Bonanos:2009wy}
and (\ref{Maxsuperalgebra-1}-\ref{Maxsuperalgebra}))
with $B=0$ and removed generator $B_C$.
It should be added that if we choose $\A=0,\B=1$ the similar contraction formulae were considered recently in \cite{biwroja10}.
\subsection{Contraction of $ OSp_R(2; 2|C)\oplus Sp(4)\quad (k=0$ case)}
For completeness we present the exotic version of Maxwell superalgebra
obtained by the contraction of \bref{wroja1.3} for $k=0$.
The bosonic algebra $Sp(4)=O(3,2)$ is given by relation \bref{OSp04}
and $OSp(2;2|C)$ is the complex superalgebra with bosonic subalgebra
$Sp(2|C)\oplus O(2|C)$ where $O_R(2|C)= O(2)\oplus O(1,1)$.
Fermionic sector of $ OSp_R(2; 2|C)$ is described by the complex $O(2)$ doublet of supercharges ${ S}^i_{+\A}$ as follows
\bea
\{{ S}^i_{+\A},{ S}^j_{+\B}\}&=&4i\,\D^{ij}(C\G_+^{\ba\rho})_{\A\B}\,\bJ_{\ba\rho}+2\,(C P_+)_{\A\B}\,\ep^{ij} \bT_+,
\nn\\
\left[\bJ_{\ba\mu},{ S}^i_{+\A}\right]&=&\frac12\,
{{ S}^i_{+\B}\,({\Gamma_{+\ba\mu}})^\B}_\A,
\qquad
\left[\bT_+,{ S}^i_{+\A}\right]=i\,\ep^{ij}\,{ S}^{j}_{+\A},
\label{OSpCp}\eea
where $\bT_+$ is the complex $O(2|C)$ generator.
The fermionic sector of complex conjugate generators $(\bJ^\dagger, { S}^{i\dagger}_+, \bT^{\dagger}_+)=
(\bJ^\dagger, { S}^{i}_-, \bT_-)$ of $\,{\overline{OSp(2; 2|C)}}$ satisfy the conjugate relations
\bea
\{{S}^i_{-\A},{ S}^j_{-\B}\}&=&-4i\,\D^{ij}(C\G_-^{\ba\rho})_{\A\B}\,\bJ^\dagger_{\ba\rho}+2\,(C\,P_-)_{\A\B}\,\ep^{ij}\, \bT_-,
\nn\\
\left[\bJ^\dagger_{\ba\mu},{ S}^i_{-\A}\right]&=&-\frac12\,
{{ S}^i_{-\B}\,({\Gamma_{-\ba\mu}})^\B}_\A,\qquad
\left[ \bT_-,{ S}^i_{-\A}\right]=i\,\ep^{ij}\,{ S}^{j}_{-\A}.
\label{OSpCm}\eea
The generators $(\bJ,{ S}^i_+, \bT_+)$ and
$(\bJ^\dagger, { S}^i_-, \bT_-)$ are commuting. The realification of $OSp(2;2|C)$ is the real superalgebra $OSp_R(2;2|C)$ extending the Lorentz algebra \bref{O31} as follows
\bea
\{{\CS}_{\A}^i,{\CS}_{\B}^j\}&=&-\,\D^{ij}\,(C\gamma^{{\mu}{\nu}})_{\A\B}
\CJ_{{\mu}{\nu}}+\ep^{ij}\,(C(\CT_0+\gamma_5\CT_5))_{\A\B},
\nn\\
\left[\CJ_{{\mu}{\nu}},{ S}_{\A}^i\right]&=&-\frac{i}2(S^i \gamma_{{\mu}{\nu}})_{\A},
\quad
\left[ \CT_0,{\CS}_{\A}^i\right]={i}\ep^{ij}\,{\CS}^j_{\A},\quad
\left[ \CT_5,{\CS}_{\A}^i\right]={i}\ep^{ij}\,({\CS}^i\gamma_5)_{\A},
\label{OSp22C7}\eea
where $
{\CS}_\A^i={S}_{+\A}^i+{S}_{-\A}^i$ and $ \bT_+=\CT_0+i\,\CT_5.$
Let us introduce the rescaled generators by the redefinitions
\bref{relgenSM3Rplus-1} and
\bea
{\bf S}_\A^i&=&\frac1{R} S_\A^i,
\label{geredefba}\\
\bT_0&=&\frac1{R^{c_0}} \CT_0,\qquad \bT_5=\frac1{R^{c_5}} \CT_5.
\label{geredefb}
\eea
In terms of the generators defined by \bref{relgenSM3Rplus-1} and
(\ref{geredefba}-\ref{geredefb})
the superalgebra $OSp_R(2;2|C)\oplus Sp(4)$ is described by the formulae \bref{BMaxalgeb02} supplemented by the following relations;
\bea
\{{\bS^{i}}_\A,{\bS^{j}}_\B\}
&=&-\,\D^{ij}\,M^2\,(C\gamma^{\mu\nu})_{\A\B}Z_{\mu\nu}
-\,\frac{\A}{R^2}\D^{ij}\,(C\gamma^{\mu\nu})_{\A\B}M_{\mu\nu}
\nn\\&&+\ep^{ij}(\frac1{R^{2-c_0}}C_{\A\B}\bT_0-\frac1{R^{2-c_5}}\,(C\gam_5)_{\A\B}\,\bT_5),\label{eq3.281}\\
\left[M_{\mu\nu},\bS^{i}_\A\right]&=&-\frac{i}{2}\,
(\bS^{i}\gam_{\mu\nu})_\A,\qquad
\left[Z_{\mu\nu},\bS^{i}_\A\right]=-\frac{i\B}{2M^2R^2}\,
(\bS^{i}\gam_{\mu\nu})_\A,\nn\\
\left[\bT_0,\bS^{i}_\A\right]&=&\frac{i}{R^{c_0}}
\ep^{ij} \bS^{j}_\A,\qquad
\left[\bT_5,\bS^{i}_\A\right]=\frac{i}{R^{c_5}}
\ep^{ij} (\bS^{j}\gamma_5)_{\A}.\label{eq3.53}\eea
If $c_0=c_5 = 2$ one gets in the contraction limit $R\to\infty$ the N=1 exotic version of Maxwell superalgebra with the basic
superalgebra relation having the form \bref{wroja1.5} obtained from \bref{eq3.281}, and includes nonvanishing
central charges $\bT_0$ and $\bT_5$. By choosing
$2> c_0, c_5 >0$
these generators are decoupled in the contraction limit, and for
$c_0=c_5=0$ we obtain $\bT_0$ and $\bT_5$ as describing the Abelian $O(2)$
symmetry generators.
\vs
\section{$N$-extended Maxwell superalgebras}
In previous section we did see that the fermionic sector (${\bf Q}_\alpha, {\bf \Sigma}_\alpha$) of simple Maxwell algebra can be obtained by the contractions from the pair of $OSp(2;4)$ supercharges
${ Q}^{i}_{\alpha}$ ($i=1,2$) or from the pair of supercharges
${ Q}_{\alpha} \in OSp(1;4)$ and ${ S}_\alpha \in OSp_R(1,2|C)$. It is interesting to see what variety of contractions can be obtained if $N>1$, i.e. for the superalgebras \bref{wroja1.3} with $2N$ Weyl supercharges.
The extension to $N>1$ of two types of contraction presented in Sect.2 is based on the observation that the standard Poincar\'{e} supercharges
satisfying relation \bref{wroja1.4} and belonging to Maxwell superalgebra
can be obtained by contraction in two ways
\begin{description}
\item[i)] from $n$ pairs of $OSp(2n;4)$ supercharges (see \bref{relgenSM3Rplus}) one gets ${\bf Q}^{i}_{\alpha}$ with $i=1\ldots n$,
\item[ii)] from $m$ pairs of supercharges of the superalgebra
${OSp_R(m;2|C)}\oplus OSp(m;4)$ (see \bref{comgeb}) providing $m$
Poincar\'e supercharges ${\bf Q}^i_{\alpha},\;(i=1\ldots m)$.
\end{description}
In general case we consider the contractions of ${OSp_R(r;2|C)}\oplus OSp(k;4)$ with $k+r=2N$.
The first type {\bf i)} of contraction describes all Poincar\'{e} superalgebra generators only
if $k=2N$ - one gets $n=N$. The second way {\bf ii)} of contracting describes the whole Poincar\'{e} supercharges sector only if $k=N$ - again we obtain $N$ Poincar\'{e} supercharges.
If $k\neq 0, N,2N$ one should apply for supercharges both types of contractions {\bf i)} and {\bf ii)}.
We shall distinguish the following two separate cases;
\begin{description}
\item[a)] $2N \geq k \geq N\geq r$ where $k+r=2N$.
\\
The supercharges in the superalgebra \bref{wroja1.3} can be described by
$({ Q}^{i}_{\alpha}, { Q}^{i'}_{\alpha}, S^{i}_{\alpha}
, (i=1,...,r,\, i'=r+1,...,k).$
From the $r$ pairs $({ Q}^{i}_{\alpha}, { S}^{i}_{\alpha})\;$
one obtains $r$ Poincar\'{e} supercharges via the second mechanism
{\bf ii)};
remaining even number $k-r=2(k-N)$ of supercharges
${ Q}^{i'}$ produce $k-N$ Poincar\'{e} supercharges as in {\bf i)}.
Concluding, we obtain in the contraction limit $r+(k-N)=N$ Poincar\'{e} supercharges satisfying the relation \bref{wroja1.4}.
\item[b)] $2N \geq r \geq N\geq k$ where $k+r=2N$.
\\
We get the supercharges (${ Q}^{i}_{\alpha}, { S}^{i}_{\alpha},
{S}^{i''}_{\alpha}),\; (i=1,...,k,\, i''=k+1,...,r)$
and we can obtain
from the $k$ pairs $({ Q}^{i}_{\alpha}, { S}^{i}_{\alpha}), (i=1,...,k),$
only $k$ Poincar\'{e} supercharges by the second mechanism (see {\bf ii)}).
The remaining $r-k=2(N-k)$ supercharges lead after contraction to the
fermionic
sector with exotic supercharges, satisfying the relation \bref{wroja1.5}.
\end{description}
The case $k=N$ is described equally well by both cases {\bf a)} and/or {\bf b)}.
\subsection{The superalgebras $OSp(k;4)$ and ${OSp_R(r;2|C)}$ }
The $OSp(k,4)$ superalgebra extends supersymmetrically
the $O(3,2)$ algebra \bref{OSp04} as follows
\bea
\{{\CQ}_{\A}^{ i},{\CQ}_{\B}^{ j}\}&=&-\,\D^{ij}\,(C\Gam^{{\8\mu}{\8\nu}})_{\A\B}
\CM_{{\8\mu}{\8\nu}}+\,2\,(C)_{\A\B}\CB^{ij},
\nn\\
\left[\CM_{{\8\mu}{\8\nu}},\CQ_{\A}^{ i}\right]&=&-\frac{i}2(\CQ^i \Gam_{{\8\mu}{\8\nu}})_{\A},\qquad
\left[\CB^{ij},{\CQ}_{\A}^{\l}\right]=-i\,\D^{\l[j}\,\CQ_{\A}^{i]},
\nn\\
\left[\CB^{ij},{\CB}^{\l m}\right]&=&-i\,\D^{\l[j}\,\CB^{i]m}
+i\,\D^{m[j}\,\CB^{i]\l}.
\label{4OSp24}\eea
The supercharges ${\CQ}_{\A}^{i},\,(\A=1,2,3,4,\,i=1,2,...,k)$ are real $SO(3,2)$ spinors and $\CB^{ij}=-\CB^{ji}\, (i,j=1,2,...,k)$ are $O(k)$ generators.
After 4+1 decomposition of the Lorentz indices $\8\mu=(\mu,4)$ the superalgebra \bref{4OSp24} can be written as $D=4$ super-AdS algebra, where $\CP_\mu=\CM_{\mu 4}$
\bea
\left[\CM_{{\mu}{\nu}},\CM_{\rho\s}\right]&=&
-i\,\h_{{\rho}[\nu}\CM_{{\mu]}\s}+i\,
\h_{{\s}[\nu}\CM_{{\mu]}\rho]},
\nn\\
\left[\CP_{{\mu}},\CM_{\rho\s}\right]&=&
-i\,\h_{{\mu}[\rho}\CP_{\s]},\qquad
\left[\CP_{{\mu}},\CP_{\nu}\right]=
-i\,\CM_{\mu\nu},
\nn\\
\{{\CQ}_{\A}^{ i},{\CQ}_{\B}^{ j}\}&=&2\,\D^{ij}\,(C\gam^{\mu})_{\A\B}
\CP_{{\mu}} \,-\D^{ij}\,(C\gam^{{\mu}{\nu}})_{\A\B}\CM_{{\mu}{\nu}}\,
+2\,(C)_{\A\B}\CB^{ij},
\nn\\
\left[\CM_{{\mu}{\nu}},\CQ_{\A}^{ i}\right]&=&-\frac{i}2(\CQ^i \gam_{{\mu}{\nu}})_{\A},\qquad
\left[\CP_{{\mu}},\CQ_{\A}^{ i}\right]=-\frac{i}2(\CQ^i \gam_{{\mu}})_{\A},
\nn\\
\left[\CB^{ij},{\CQ}_{\A}^{\l}\right]&=&-i\,\D^{\l[j}\,\CQ_{\A}^{ i]},\qquad
\left[\CB^{ij},{\CB}^{\l m}\right]=-i\,\D^{\l[j}\,\CB^{i]m}
+i\,\D^{m[j}\,\CB^{i]\l}.
\label{OSp242}\eea
The $ OSp(r; 2|C)$ superalgebra has the following form
\bea
\left[\bJ_{\ba\mu},\bJ_{\ba\nu}\right]&=&{i}\,{\ep_{\ba\mu\ba\nu}}^{\ba\rho}
\bJ_{\ba\rho},\qquad
\left[{\bf T}_+^{ij},{\bf T}_+^{\l m}\right]=-i\,(\D^{\l[j}\,{\bf T}_+^{i]m}-
\D^{m[j}\,{\bf T}_+^{i]\l}),
\nn\\
\{{S}^i_{+\A},{S}^j_{+\B}\}&=& 2i \left(2\,\D^{ij}(C\G_+^{\ba\rho})_{\A\B}\,\bJ_{\ba\rho}-i
\,(C P_+)_{\A\B}\,{\bf T}_+^{ij}\right),
\nn\\
\left[\bJ_{\ba\mu},{S}^i_{+\A}\right]&=&\frac12\,
{S^i_{+\B}\,({\Gamma_{+\ba\mu}})^\B}_\A,\qquad
\left[{\bf T}_+^{ij},{S}^{\l}_{+\A}\right]=-i\,\D^{{\l}[j}\,S^{i]}_{+\A},
\label{OSpNCp}\eea
where $\bJ_{\ba\mu},(\ba\mu=0,1,2)$, ${\bf T}_+^{ij}, (i=1,...,r)$ and
$S^i_{+\A}$ are complex $SO(2,1)$, $O(r)$ and supersymmetry
generators respectively.
The complex conjugate generators
$(\bJ, S^{i}_+,{\bf T}_+)^{\dagger}=
(\bJ^\dagger, S^{i}_-,{\bf T}_-)$ describing the superalgebra
${\overline{OSp({r},2|C)}}$ satisfying the conjugate relations (for $r=2$
see also \bref{OSpCm}).
One obtain finally the following real $OSp_R(r;2|C) $ superalgebra
\bea
\left[\CJ_{{\mu}{\nu}},\CJ_{\rho\s}\right]&=&-i\,
\h_{{\rho}[\nu}\CJ_{{\mu]}\s}+i\,\h_{{\s}[\nu}\CJ_{{\mu]}\rho},
\qquad
\left[\CBA^{ij},\CBA^{k{\ell}}\right]=-{i}(\D^{k[j}\,\CBA^{i]{\ell}}-
\D^{\l[j}\,\CBA^{i]k}),
\nn\\
\left[\CBA^{ij},\CBB^{k{\ell}}\right]&=&-{i}(\D^{k[j}\,\CBB^{i]{\ell}}-
\D^{{\ell}[j}\,\CBB^{i]k}),\qquad
\left[\CBB^{ij},\CBB^{k{\ell}}\right]={i}(\D^{k[j}\,\CBA^{i]{\ell}}-
\D^{\l[j}\,\CBA^{i]k}),\nn\\
\{{\CS}_{\A}^i,{\CS}_{\B}^j\}&=&-\,\D^{ij}\,(C\Gam^{{\mu}{\nu}})_{\A\B}\CJ_{{\mu}{\nu}}+\,(C({\CBA}^{ij}-\gamma_5\CBB^{ij}))_{\A\B},\label{OSp22C8}\\
\left[\CJ_{{\mu}{\nu}},\CS_{\A}^i\right]&=&-\frac{i}2(\CS^i \gam_{{\mu}{\nu}})_{\A},\quad
\left[\CBA^{ij},\CS_{\A}^k\right]=-{i}\D^{k[j}\,\CS^{i]}_{\A},\quad
\left[\CBB^{ij},\CS_{\A}^k\right]=-{i}\D^{k[j}\,(\CS^{i]}\gamma_5)_{\A},
\nn\\ \label{OSp2CRalg}\eea
where the real generators $\CJ_{\mu\nu}, \CS^i_\A, \CBA^{ij}, \CBB^{ij}$ describe respectively
$SO(3,1)$, supersymmetry and the pair of internal $SO(r)$ generators that are related to the complex ones by the formula \bref{bJdef} and the relations
$
{S}_{+\A}^i+{S}_{-\A}^i=\CS_\A^i,\, $ $
{\bf T}_+^{ij}=\frac12(\CBA^{ij}+i\,\CBB^{ij}).$
\subsection{Contraction of $OSp_R(r;2|C)\oplus OSp(k;4),\; (k\geq r\geq 0)$}
In this subsection we consider the case {\bf a)} $(k\geq r)$.
In order to obtain after contraction $R\to\infty$ the Maxwell algebra as bosonic subalgebra we rescale the generators $(\CM_{\mu\nu},\CP_\mu,\CJ_{\mu\nu})\in O(3,2)\oplus O(3,1)$ in accordance with the formulae \bref{relgenSM3Rplus-1} .
In the contraction limit $(R\to\infty)$ we obtain the bosonic Maxwell
algebra \bref{wroja1b} for any value of $(\A,\B=1-\A)$.
The generators of the internal symmetries $ O(r|C)\oplus O(k), \;(k\geq r)$
are split into three families,
\bea
1) &&(\CB^{ij},\CT_0^{ij},\CT_5^{ij})\in O(r)\oplus O_R(r|C), \quad (i,j=1,...,r),\; \nn\\
2) && \CB^{i'j'} \in O(k-r)=O(2(k-N)), \quad (i',j'=r+1,...,k),\quad
\label{intgene}\\
3) && \CB^{ij'} \in O(k)/(O(r)\oplus O(k-r)), \quad (i=1,...,r,\, j'=r+1,...,
k).\nn\eea
In the first group of the generators \bref{intgene} the diagonal subalgebra $O(r)$ remain unscaled
\be \bB_D^{ij}=\CB^{ij}+\CT_0^{ij} \label{eq3.9}\ee
while the remaining ones are rescaled as \footnote{
In general case one can consider the rescaling
$\bT_0^{ij}=\frac1R((1-\A')\,\CT_0^{ij}-\A'\,\CB^{ij})$. It appears however that the contraction limit $R\to\infty$ will not depend on $\A'$, so we choose for simplicity $\A'=0$.}
\be
\bT_0^{ij}=\frac1R\,\CT_0^{ij},\qquad
\bT_5^{ij}=\frac1R\,\CT_5^{ij}.
\label{BTrescaleb} \ee
The generators $\CB^{i'j'} =\CB^{i'j'}_++\CB^{i'j'}_-\in O(k-r)$ we
decompose as follows
\be
\CB_-^{i'j'}\in U(k-N),\quad
\CB_+^{i'j'}\in O(2(k-N))/U(k-N),
\label{defBpm}\ee
where $2(k-r)\times 2(k-r)$ matrix of generators $\CB=(\CB^{i'j'})$ is decomposed as follows
\be
\CB
\begin{pmatrix}A_0 & S \\ -S & A_0 \end{pmatrix}+\begin{pmatrix}
A_3 & A_1 \\ A_1 & -A_3 \end{pmatrix}=\CB_-+\CB_+,
\ee
where the matrices $S$ and $A_\l\, (\l=1,2,3)$ satisfy $S^T=S$ and $A_\l^T=-A_\l$.
If we introduce the anti-symmetric matrix $\W=\begin{pmatrix}
0 & 1_{k-N} \\ -1_{k-N} & 0 \end{pmatrix}$, the matrices $\CB_\mp$ satisfy the
relations $ \W\, \CB_\pm \pm \CB_\pm\,\W=0. $
The generators $\CB^{i'j'}_\mp $ we rescale as follows
\be
\bB^{i'j'}_-=\,\CB^{i'j'}_-,\qquad
\bB^{i'j'}_+=\frac1R\,\CB^{i'j'}_+.
\label{eq3.14}\ee
The remaining internal symmetry generators $\CB^{ij'}$ which occur in the anti-commutators of $\{\bQ^i,\bQ^{j'}\}$ and $\{\bQ^i,\bSig^{j'}\}$ are rescaled by
\be
\bB^{ij'}=\frac1R\,\CB^{ij'}.
\label{eq3.15}\ee
If we perform the contraction $R\to\infty$ of the internal symmetry generators listed in \bref{intgene} we obtain the pair of non-Abelian Lie algebras $O(r)$ (generators $\bB_D^{ij})$ and $U(N-r)$ (generators $\bB_-^{i'j'})$; remaining contracted generators $\bT_0^{ij}, \bT_5^{ij}, \bB_+^{i'j'}, \bB^{ij'}$ are becoming Abelian.
The $k$ supercharges $\CQ^i,(i=1,...,k)$ are split into $r$ supercharges $\CQ^i,(i=1,...,r)$ and remaining $k-r$ supercharges $\CQ^{i'},(i'=r+1,...,k)$.
The first ones are combined with $r$ supercharges $\CS^i,(i=1,...,r)$
to define
\be
\bQ^i_\A=\frac1{R^{1/2}}\,(\CQ^i_\A +\CS^i_\A ) ,\qquad
\bSig^i_\A =\frac1{R^{3/2}} \,\CS^i_\A.
\label{eq3.16}\ee
The remaining $\CQ^{i'},(i'=r+1,...,k)$ are used in order to define
\be
{\bQ^{i'}_\A }= \frac{1}{R^{1/2}}\,\CQ^{j'}_\B \,{{\Pi^+}^{j'i'\B}}_{\A},
\qquad
{\bSig^{i'}_\A }=\frac{1}{R^{3/2}}\,\CQ^{j'}_{\B}\,{{\Pi^-}^{j'i'\B}}_{\A} ,
\label{eq3.17}\ee
where $\Pi^\pm$ are the projection operators ( we recall that $\gamma_5^2=\W^2=-1$),
\be
{{\Pi^\pm}^{j'i'\B}}_{\A}=\frac12({{\D}^{j'i'}}{\D^\B}_{\A}\pm{{\gamma_5}^\B}_\A
{{\W}^{j'i'}}),\qquad \Pi^\pm \Pi^\pm=\Pi^\pm,\quad \Pi^\pm \Pi^\mp=0.
\ee
Supercharges $\bQ^{i'},\bSig^{i'}, (i'=r+1,...,k)$ satisfy
the identities
\be \bQ^{i'}=\bQ^{j'}\,\Pi^{+j'i'},\qquad \bSig^{i'}=\bSig^{j'}\,\Pi^{-j'i'},
\ee
but because $k-r=2(k-N)$ the number of independent supercharge components is reduced to $k-N$ for both $\bQ^{i'}$ and $\bSig^{i'}$.
One can write down the superalgebra $OSp_R(r;2|C)\oplus OSp(k;4)$
in terms of rescaled generators $(P_\mu,M_{\mu\nu},Z_{\mu\nu},\bQ^i,\bQ^{i'},\bSig^i,\bSig^{i'},\bB_D^{ij},\bT_0^{ij},\bT_5^{ij},\bB_-^{i'j'},\bB_+^{i'j'},\bB^{ij'})$, (see \bref{relgenSM3Rplus-1}, (\ref{eq3.9}-\ref{BTrescaleb}), (\ref{eq3.14}-\ref{eq3.17})). If we perform the contraction limit $R\to\infty$ we obtain the following relations describing $N$-extended $(N>1)$ Maxwell superalgebras,
\bea
\left[P_\mu,P_{\nu}\right]&=&i\,M^2\,Z_{\mu\nu},\qquad
\left[M_{\mu\nu},M_{\rho\s}\right]=-i\,\h_{\rho[\nu}M_{\mu] \s}+i\,
\h_{\s[\nu}M_{\mu]\rho},\nn\\
\left[P_\mu,M_{\rho\s}\right]&=&-i\,\h_{\mu[\rho}P_{\s]},\qquad
\left[M_{\mu\nu},Z_{\rho\s}\right]=-i\,\h_{\rho[\nu}Z_{\mu] \s}+i\,
\h_{\s[\nu}Z_{\mu]\rho},
\eea
\bea
\left[P_\mu,\bQ^i\right]&=&\frac{i}{2}\,\bSig^i\,\gam_\mu,\quad
\left[M_{\mu\nu},\bQ^i\right]=-\frac{i}2\,\bQ^i\gam_{\mu\nu},\quad
\left[M_{\mu\nu},\bSig^i\right]=-\frac{i}2\,\bSig^i\gam_{\mu\nu},
\nn\\
\left[P_\mu,\bQ^{i'}\right]&=&-\frac{i}{2}\,\bSig^{i'}\,\gam_\mu,\quad
\left[M_{\mu\nu},\bQ^{i'}\right]=-\frac{i}2\,\bQ^{i'}\gam_{\mu\nu},\quad
\left[M_{\mu\nu},\bSig^{i'}\right]=-\frac{i}2\,\bSig^{i'}\gam_{\mu\nu},
\nn\\
\left[Z_{\mu\nu},\bQ^{i}\right]&=&
\left[Z_{\mu\nu},\bQ^{i'}\right]=
\left[Z_{\mu\nu},\bSig^{i}\right]=\left[Z_{\mu\nu},\bSig^{i'}\right]=0, \eea
\bea
\{{\bQ^i}_\A,{\bQ^j}_\B\}&=&
2\,\D^{ij}\,(C\gamma^{\mu})_{\A\B}P_\mu
-\,C_{\A\B}\,\bT_0^{ij}-\,(C\gam_5)_{\A\B}\,\bT_5^{ij},\quad
\nn\\
\{{\bQ^i}_\A,{\bSig^j}_\B\}&=&
-\,{\D^{ij}}\,(C\gamma^{\mu\nu})_{\A\B}M^2\,Z_{\mu\nu},
\nn\\
\{{\bQ^i}_\A,{\bQ^{j'}}_\B\}&=&{2}\,(C\bB^{i,\l'}\Pi^{+\l'j'})_{\A\B}.
\nn\\
\{{\bQ^{i'}}_\A,{\bQ^{j'}}_\B\}&=&2(C\gam^\mu\Pi^+)_{\A\B}^{i'j'}\,P_\mu
+\,{2}\,(C\bB_+\Pi^+)^{i',j'}_{\A\B},
\nn\\
\{{\bQ^{i'}}_\A,{\bSig^{j'}}_\B\}&=&(C\gam^{\mu\nu}\Pi^-)_{\A\B}^{i'j'}\,M^2\,Z_{{\mu}{\nu}}.\label{QQalg}\eea
The internal symmetry sector containing the generators $\bB_D^{ij}\in O(k)$
and
$\bB_-^{i'j'}\in U(N-k)$ is described by the following non-vanishing commutators,
\bea
\left[\bB_D^{ij},\bQ^\ell_\A\right]&=&-i\D^{\ell[j} \bQ^{i]}_{\A},\qquad
\left[\bB_D^{ij},\bSig^\ell_\A\right]=-i\D^{\ell[j} \bSig^{i]}_{\A},
\nn\\
\left[\bB_D^{ij},\bB_D^{\ell m}\right]&=&-i(\D^{\ell[j} \bB_D^{i]m}-
\D^{m[j} \bB_D^{i]\ell}),\qquad
\left[\bB_D^{ij},{\bT_0}^{\ell m}\right]=-i(\D^{\ell[j} \bT_0^{i]m}-
\D^{m[j} \bT_0^{i]\ell}),
\nn\\
\left[\bB_D^{ij},\bT_5^{\ell m}\right]&=&-i(\D^{\ell[j} \bT_5^{i],m}-
\D^{m[j} \bT_5^{i]\ell})
,\qquad
\left[\bB_D^{ij},\bB^{\ell m'}\right]=-i\D^{\ell[j} \bB^{i]m'},
\eea
\bea
\left[\bT_0^{ij},\bQ^\ell_\A\right]&=&-i\D^{\ell[j} \bSig^{i]}_{\A},
\quad
\left[\bT_5^{ij},\bQ^\ell_\A\right]=-i\D^{\ell[j} \bSig^{i]}_{\A}\gamma_5,
\quad
\left[\bB^{ij'},\bQ^\ell_\A\right]={i}\D^{i\ell} \bSig^{j'}_\A,
\nn\\
\left[\bB^{ij'},\bQ^{\ell'}_\A\right]&=& {i}( \bSig^{i}\Pi^{+j'\l'})_\A,
\quad
\left[\bB_-^{i'j'},\bQ^{\ell'}_\A\right]=
-i\bQ^{[i'}\,\Pi^{+j']\l'}, \quad
\left[\bB_+^{i'j'},\bQ^{\ell'}_\A\right]
=-i\bSig^{[i'}\,\Pi^{+j']\l'}, \nn\\
\left[\bB_-^{i'j'},\bSig^{\ell'}_\A\right]&=&-i\bSig^{[i'}\,\Pi^{-j']\l'},
\quad
\left[\bB_-^{i'j'},\bB^{\ell m'}\right]=
{i}\,\rho^{ij,mn}_-\,\bB^{\l n'}.\eea
We see from \bref{QQalg} that all fermionic generators $\bQ^i_\A$ and $\bQ^{i'}_\A$ are the Poincar\'e supercharges.
\subsection{Contraction of $OSp_R(r;2|C)\oplus OSp(k;4),\; (r\geq k\geq 0)$}
The generators of the internal symmetries $ O(r|C)\oplus O(k), \;(k\leq r)$
are split into the following three families,
\bea
1) &&(\CB^{ij},\CT_0^{ij},\CT_5^{ij})\in O(k)\oplus O_R(k|C), \quad (i,j=1,...,k),\; \nn \\
2) && (\CT_0^{i''j''}, \CT_5^{i''j''}) \in O_R(r-k|C)=O_R(2(r-N)|C), \quad (i'',j''=k+1,...,r),\quad \nn \\
3) && (\CT_0^{ij''}, \CT_5^{ij''})\in O_R(r|C)/(O_R(k|C)\oplus O_R(r-k|C)),
\quad (i=1,...,k,\, j''=k+1,...,r).\nn \eea
In the first group we rescale the generators $\bB^{ij},\bT_0^{ij},\bT_5^{ij}$
as in previous subsection (see (\ref{eq3.9}-\ref{BTrescaleb})).
The generators
$ (\CT_0^{i''j''}, \CT_5^{i''j''}) $ and $(\CT_0^{ij''}, \CT_5^{ij''})$ are rescaled as follows
\be
\bT_0^{i''j''}=\frac1{R^2}\CT_0^{i''j''},\quad
\bT_5^{i''j''}=\frac1{R^2}\CT_5^{i''j''},\quad
\bT_0^{ij''}=\frac1{R^{3/2}}\CT_0^{ij''},\quad
\bT_5^{ij''}=\frac1{R^{3/2}}\CT_5^{ij''}.
\ee
After the contraction limit $R\to\infty$ only the generators $\bB_D^{ij}$ describe the non-Abelian $O(k)$ generators; remaining generators $ (\bT_0^{ij},
\bT_5^{ij},\bT_0^{ij''}, \bT_5^{ij''},\bT_0^{i''j''}, \bT_5^{i''j''}) $
are becoming Abelian.
The $r$ supercharges $\CS^i,(i=1,...,r)$ are split into $k$ supercharges
$\CS^i,(i=1,...,k)$ and remaining $r-k$ supercharges $\CS^{i''},(i''=k+1,...,r)$.
The first ones are combined with $k$ supercharges $\CQ^i,(i=1,...,k)$
to define
\be
\bQ^i_\A=\frac1{R^{1/2}}\,(\CQ^i_\A +\CS^i_\A ) , \qquad
\bSig^i_\A = \frac1{R^{3/2}} \,\CS^i_\A .
\ee
The remaining generators $\CS^{i''},(i''=k+1,...,r)$ are rescaled as
\bea
\bS^{i''}=\frac1{R}\,\CS^{i''}.
\eea
In the contraction limit $R\to\infty$ we obtain besides the Maxwell algebra
\bref{wroja1b} the following set of algebraic relations
\bea
\left[M_{\mu\nu},\bQ^i\right]&=&-\frac{i}2\,\bQ^i\gam_{\mu\nu},\quad
\left[M_{\mu\nu},\bSig^i\right]=-\frac{i}2\,\bSig^i\gam_{\mu\nu},\quad
\left[M_{\mu\nu},\bS^{i''}\right]=-\frac{i}2\,\bS^{i''}\gam_{\mu\nu},
\nn\\
\left[P_\mu,\bQ^i\right]&=&\frac{i}{2}\,\bSig^i\,\gam_\mu,\quad
\left[\bT_0^{ij},\bQ^\ell_\A\right]=-i\D^{\ell[j} \bSig^{i]}_{\A},\quad
\left[\bT_5^{ij},\bQ^\ell_\A\right]=-i\D^{\ell[j} \bSig^{i]}_{\A}\gam_5.
\eea
\bea
\{{\bQ^i}_\A,{\bQ^j}_\B\}&=&
2\,\D^{ij}\,(C\gamma^{\mu})_{\A\B}P_\mu
-\,C_{\A\B}\,\bT_0^{ij}-\,(C\gam_5)_{\A\B}\,\bT_5^{ij},\quad
\nn\\
\{{\bQ^i}_\A,{\bSig^j}_\B\}&=&
-\,{\D^{ij}}\,(C\gamma^{\mu\nu})_{\A\B}\,M^2\,Z_{\mu\nu},
\nn\\
\{{\bQ^i}_\A,\bS^{j''}_\B\}&=&
(C(\bT_0^{ij''}-\gam_5 \bT_5^{ij''}))_{\A\B},
\nn\\
\{{\bS^{i''}}_\A,{\bS^{j''}}_\B\}
&=&-\,\D^{i''j''}\,(C\gamma^{\mu\nu})_{\A\B}\,M^2\,Z_{\mu\nu}
+(C_{\A\B}\bT_0^{i''j''}-\,(C\gam_5)_{\A\B}\,\bT_5^{i''j''}).
\label{eq3.32}\eea
\bea
\left[\bB_D^{ij},\bQ^\ell_\A\right]&=&-i\D^{\ell[j} \bQ^{i]}_{\A},\quad
\left[\bB_D^{ij},\bSig^\ell_\A\right]=-i\D^{\ell[j} \bSig^{i]}_{\A},
\quad
\left[\bB_D^{ij},\bB_D^{\ell m}\right]=-i(\D^{\ell[j} \bB_D^{i]m}-
\D^{m[j} \bB_D^{i]\ell}),\nn\\
\left[\bB_D^{ij},{\bT_0}^{\ell m}\right]&=&-i(\D^{\ell[j} \bT_0^{i]m}-
\D^{m[j} \bT_0^{i]\ell}),\quad
\left[\bB_D^{ij},\bT_5^{\ell m}\right]=-i(\D^{\ell[j} \bT_5^{i],m}-
\D^{m[j} \bT_5^{i]\ell})
,\qquad
\nn\\
\left[\bB_D^{ij},\bT_0^{\ell m''}\right]&=&-i\D^{\ell[j} \bT_0^{i]m''},
\qquad
\left[\bB_D^{ij},\bT_5^{\ell m''}\right]=-i\D^{\ell[j} \bT_5^{i]m''}.
\eea
We see from relations \bref{eq3.32} that the Maxwell superalgebras derived in this subsection have hybrid structure: the anti-commutators of $k$ Weyl supercharges $\bQ^i$ as in Poincar\'e SUSY algebra
describe the fourmomenta generators $P_\mu$ (see \bref{wroja1.4}), but remaining $r-k$ supercharges $\bS^{i''}$ supersymmetrizing the tensorial generators $Z_{\mu\nu}$ describe the exotic sector of $N$-extended Maxwell superalgebra with the SUSY relations \bref{wroja1.5} .
\section{Discussion}
The aim of this paper is to provide a large class of superalgebraic structures describing $N$-extended Maxwell superalgebra. In Sect.2 we considered simple Maxwell superalgebras $(N=1)$; the extended case $(N>1)$ is described in Sect.3.
For arbitrary $N$ we can considered the contractions of $2N$ different superalgebras described by \bref{wroja1.3} with $k=1,2,...,2N$. The case $k=0$ as it was argued in introduction and in subsections 2.4 and 3.3 provides only the purely exotic supersymmetrization of Maxwell algebra with all supercharges satisfying the relations \bref{wroja1.5}. If $N\leq k\leq 2N$ we obtain the Maxwell superalgebra with $N$ "standard" SUSY relations \bref{wroja1.4}; if $0\leq k<N$ one obtains a hybrid Maxwell superalgebra, with $k$ standard Poincar\'{e} supercharges and $r-k$ supercharges satisfying exotic SUSY relation \bref{wroja1.5}.
At present it is not known whether the exotic SUSY relation \bref{wroja1.5} can provide some dynamical models with physically interesting consequences
but for completeness of our algebraic considerations we included as well the extended Maxwell superalgebras with exotic supercharges.
The contractions considered in this paper can be further applied to the description of various Maxwell supergravity models. For example in order to obtain $N=1$ Maxwell supergravity one should consider firstly the gauge fields on $O(3,1)\oplus OSp(2,4)$ or $OSp_R(1,2|C)\oplus OSp(1,4)$ superalgebras and then introduce the gauge formulations of deformed Maxwell supergravity models.
In the following step one can look for the deformed Maxwell supergravity
actions with finite contraction limits
defined in accordance with the contraction prescription in Sect.2.2 and 2.3. In general case one can consider the gauge theories on superalgebras \bref{wroja1.3} and study by their suitable contractions the possible new extended Maxwell supergravity models.
\vs
{\bf Acknowledgements }
The authors would like to thank Jose A.~de Azc\'{a}rraga, Joaquim Gomis and Piotr Kosinski for their valuable remarks and discussions. The paper has been supported by Polish NCN grant No.2011/ST2/03354.
|
\section{Introduction}
It is well known that the superfield formulations of supersymmetric
field theories, with the maximal number of the underlying
supersymmetries being manifest and off-shell, are extremely useful
for studying quantum aspects of these theories. In many cases, such
formulations not only drastically reduce the amount of perturbative
calculations, but also allow one to make certain conjectures about a
possible structure of the final results prior to any calculation.
The $\cN=1$ superspace \cite{BK-book} is natural for $\cN=1$, $d=4$
supersymmetric models, while the adequate superfield approach to
$\cN=2$, $d=4$ theories is offered by $\cN=2$ harmonic superspace
\cite{hss-book}.
As for the renowned $\cN=4$ SYM theory, no appropriate formulation of it in
terms of unconstrained off-shell $\cN=4$ superfields is known to date.
On shell, the $\cN=4$ SYM theory is equivalent to the $\cN=3$ SYM theory (see \cite{hss-book} and refs.\ therein).
The latter possesses an unconstrained superfield formulation in $\cN=3$ harmonic superspace \cite{GIKOS1,GIKOS2}, such
that three out of four supersymmetries of the original theory are manifest and off-shell within this framework.
This approach proved to be very fruitful for establishing quantum finiteness of $\cN=3$ SYM theory \cite{DM}, as well as
for constructing the $\cN=3$ supersymmetric Born-Infeld theory
\cite{IZ}.\footnote{A possible scale-invariant generalization of the $\cN=3$ Born-Infeld theory was discussed in \cite{BISZ}.}
The basic goal of the present paper is to provide an evidence that the $\cN=3$ harmonic superspace approach is also
useful for studying the low-energy effective actions in the $\cN=3$ and $\cN=4$ SYM theories.
It is known that the leading terms in the $\cN=4$ SYM effective action in the $\cN=2$ superfield formulation (which manifests only
two out of four supersymmetries) are described by the non-holomorphic
potential \cite{DS,S,deWit:1996,non-hol1,non-hol2,non-hol3,non-hol4,non-hol5,review}.
The hypermultiplet completion of the non-holomorphic potential,
such that it ensures the on-shell $\cN=4$ supersymmetry, was found in
\cite{BuIv} (and further elaborated on in refs.
\cite{BIP,BBP,BuPl}). The full action contains a scale-invariant
and SU(4) symmetric $F^4/X^4$ term, as well as some other terms
related to this leading one by $\cN=4$ supersymmetry. Here
$F_{mn}$ is the Maxwell field strength and $X^2$ is the square of
SU(4) invariant norm of scalar fields.
In the present paper we develop the $\cN=3$ harmonic superspace description of the leading terms in the
$\cN=4$ SYM effective action to the order $F^4/X^4$. We seek this action as an integral over the analytic
subspace of the $\cN=3$ harmonic superspace, with the Lagrangian density being a local functional of the
analytic superfield strengths without derivatives on them. We show
that the requirements of the scale and $\gamma_5$ invariance uniquely fix
the form of this functional. We check that the action constructed respects the full SU(2,2$|$3) superconformal symmetry
and, in components, yields the $F^4/X^4$ term, where $X^2=\varphi^i\bar\varphi_i$ is the bilinear SU(3)
(and in fact SU(4)) invariant of the involved scalar fields.
We stress that the obtained action is essentially defined on the
Coulomb branch of the theory, when the scalar fields acquire
non-vanishing vevs, $c^i=\langle\varphi^i \rangle\ne0$. These
constants $c^i$ explicitly appear in the effective action, so that the
effective Lagrangian is singular at $c^i=0$. However, we show that
the action is in fact independent of any particular choice of $c^i$,
$c^i\ne0$. This is entirely analogous to what happens in the $\cN=2$
harmonic superspace formulation of the $\cN=2$ improved tensor
multiplet model given in \cite{GIO}. It was emphasized there that the
presence of such constants in the action has a topological origin.
In accord with this interpretation, the low-energy $\cN=4$
SYM effective action contains a topological term given by the
Wess-Zumino action for the scalar fields \cite{TZ,Intriligator}.
Therefore the presence of such constants in the effective
Lagrangian is not surprising.
One of the advantages of the $\cN=3$ superspace formulation is the
possibility to go off shell due to the existence of unconstrained
gauge prepotentials. Varying with respect to these prepotentials, we obtain the
effective equations of motion corresponding to the effective action.
Like in the non-scale-invariant $\cN=3$ Born-Infeld
theory \cite{IZ}, elimination of some of the auxiliary fields from the
effective equations of motion allows one to reproduce not only $F^4/X^4$ term,
but also the $F^6/X^8$ term in the effective action, which precisely
matches with that appearing in the component expansion of the conformally-invariant Born-Infeld
action. Therefore, the effective action obtained reproduces the
worldvolume action of D3 brane on the $AdS_5\times S^5$ background up to the order
$F^6/X^8$.
The paper is organized as follows. Section 2 contains a brief
summary of the basic ingredients of the $\cN=3$ harmonic
superspace formalism. In particular, we give the representation of
the $\cN=3$ superconformal group SU(2,2$|$3) on the $\cN=3$ SYM
superfield strengths. In Section 3, employing the scale and
$\gamma_5$ invariance, we derive the $\cN=3$ SYM low-energy
effective action and then show its invariance under the full
$\cN=3$ superconformal group. In Section 4 we derive the $F^4/X^4$
and $F^6/X^8$ component terms from the superfield action. The last
Section is devoted to discussing some open problems deserving
further study. In Appendices A and B we collect some technical
details concerning the derivation of the superfield action and
calculation of SU(3) harmonic integrals. A possible
four-dimensional representation of the Wess-Zumino term with
manifest SU(3) symmetry is discussed in the Appendix C. In
Appendix D we demonstrate that the $\cN=3$ superfield effective
action proposed can be used for studying the effective superfield
equations of motion.
Throughout the paper we follow the $\cN=3$ superspace
conventions employed in \cite{BISZ} and \cite{IZ}.
\section{$\cN=3$ SYM setup}
\subsection{Superfield strengths in $\cN=3$ harmonic superspace}
The standard $\cN=3$ superspace is parametrized by the coordinates
$z^M=(x^m,\theta_i^\alpha,\bar\theta^{i\dot\alpha})\,$, where
$i=1,2,3$ is the SU(3) triplet index. Following \cite{GIKOS1,GIKOS2},
we introduce the SU(3) harmonic variables
$u^I_i=(u^1_i, u^2_i, u^3_i)$ and their conjugates,
$\bar u_I^i=(\bar u_1^i, \bar u_2^i,\bar u_3^i)$, with the
properties
\be
u^I_i \bar u^i_J=\delta^I_J\,,\quad
u^I_i \bar u^j_I=\delta_i^j\,,\quad
\varepsilon^{ijk}u^1_i u^2_j u^3_k=1\,.
\ee
These defining relations are the orthogonality and
completeness conditions. The harmonic variables allow one to convert the small indices $i,j,\ldots$ on
which the R-symmetry SU(3) group is linearly realized, into the capital indices, $I,J,\ldots$, which are inert under SU(3).
For instance, we will make use of the projected Grassmann variables,
$\theta_I^\alpha=\theta_i^\alpha \bar u^i_I$,
$\bar\theta^{I\dot\alpha}=\bar\theta^{i\dot\alpha} u^I_i$. Some of
these projected Grassmann variables parametrize the analytic
subspace,
\be
\{\zeta_A, u \}=\{x_A^{m},\theta_2^\alpha,\theta_3^\alpha,
\bar\theta^{1\dot\alpha},\bar\theta^{2\dot\alpha},u \}\,,\qquad
x_A^{m}=x^m-i\theta_1\sigma^m\bar\theta^1+i\theta_3\sigma^m\bar\theta^3\,.
\label{anal-coord}
\ee
The analytic superspace (\ref{anal-coord}) is closed
under the $\cN=3$ supersymmetry \cite{GIKOS1,GIKOS2}, and, hence, plays a role similar
to that of usual chiral subspace in the $\cN=1$ superspace \cite{BK-book} and of the $\cN=2$ harmonic
analytic superspace \cite{hss-book}.
The harmonic projections of the covariant spinor derivatives\footnote{
We use the following rules of converting the vector and bi-spinorial indices into
each other,
$x_{\alpha\dot\alpha}=(\sigma^m)_{\alpha\dot\alpha}x_m$,
$x_m=\frac12(\tilde\sigma_m)^{\dot\alpha\alpha}x_{\alpha\dot\alpha}$,
$\partial_{\alpha\dot\alpha}=\frac12(\sigma^m)_{\alpha\dot\alpha}\partial_m$,
$\partial_m=(\tilde\sigma_m)^{\dot\alpha\alpha}\partial_{\alpha\dot\alpha}$.},
\be
D^i_\alpha=\frac{\partial}{\partial\theta_i^\alpha}+2i\bar\theta^{i\dot\alpha}
\frac{\partial}{\partial x^{\alpha\dot\alpha}}\,,\qquad
\bar D_{i\dot\alpha}=-\frac{\partial}{\partial\bar\theta^{i\dot\alpha}}-
2i\theta_i^\alpha\frac{\partial}{\partial x^{\alpha\dot\alpha}}\,,
\ee
are given by $D^I_\alpha=D^i_\alpha u^I_i$ and
$\bar D_{I\dot\alpha}=\bar D_{i\dot\alpha}\bar u_I^i$. It is
important that in the analytic coordinates
(\ref{anal-coord}) two of these six derivatives become short,
\be
D^1_\alpha=\frac\partial{\partial\theta_1^\alpha}\,,\qquad
\bar
D_{3\dot\alpha}=-\frac\partial{\partial\bar\theta^{3\dot\alpha}}\,,
\label{anal-deriv}
\ee
thus demonstrating that the $\cN=3$ analytic superfields (i.e.\ those living on the analytic superspace (\ref{anal-coord}))
can be covariantly defined by the Grassmann Cauchy-Riemann
conditions,
\be
D^1_\alpha \Phi(z, u) =\bar D_{3\dot\alpha} \Phi(z, u) = 0\quad
\Rightarrow \quad\Phi(z, u) = \hat\Phi(\zeta_A, u)\,.
\ee
The explicit expressions for the other four derivatives in the analytic basis can be
found in the appendix of our previous paper \cite{BISZ}, where the
harmonic derivatives $D^I_J$ are also written down.
Among these harmonic derivatives, $D^1_2$, $D^2_3$ and $D^1_3$
commute with (\ref{anal-deriv}) and so preserve the analyticity, while the remaining three $D^2_1$, $D^3_2$, $D^3_1$
do not. These six harmonic derivatives, together with the U(1)
charges $S_1$ and $S_2$, form an su(3) algebra \cite{AFSZ}.
The conventional $\cN=3$ SYM superfield strengths in the standard $\cN=3$ superspace
are described by the antisymmetric SU(3) tensor superfields
$W^{ij}=-W^{ji}$. In the linearized approximation, these superfields obey the constraints \cite{HST},
\be
D^i_\alpha W_{jl}=\frac12(\delta^i_j D^k_\alpha W_{kl}-\delta^i_l D^k_\alpha W_{kj})
\,,\qquad
\bar D_{i\dot\alpha}W_{jk}+\bar D_{j\dot\alpha}W_{ik}=0\,,
\label{constr}
\ee
which eliminate all non-physical components in these superfields and
put the physical ones on shell. Projecting these superfield
strengths on the harmonic variables, we obtain the following six
superfields,
\bea
&&
\bar W^{12}=u^1_i u^2_j\bar W^{ij}\,,\quad
\bar W^{23}=u^2_i u^3_j\bar W^{ij}\,,\quad
\bar W^{13}=u^1_i u^3_j\bar W^{ij}\,,\quad
\nn\\&&
W_{12}=\bar u_1^i \bar u_2^j W_{ij}\,,\quad
W_{23}=\bar u_2^i \bar u_3^j W_{ij}\,,\quad
W_{13}=\bar u_1^i \bar u_3^j W_{ij}\,.
\label{W_}
\eea
It is straightforward to find the harmonic projections of the
constraints (\ref{constr}), which gives rise to a number of differential
relations among the superfields (\ref{W_}).
Consider, for instance, $\bar W^{12}$ and $W_{23}$. They
obey the following (on-shell) constraints \cite{AFSZ}\footnote{The constraints (\ref{anal})--(\ref{linearity}) can
also be derived by quantizing a massless superparticle moving in the $\cN=3$
harmonic superspace \cite{BS08}.}:
\begin{itemize}
\item[(i)] First-order analyticity constraints,
\bea
&&
D^1_\alpha \bar W^{12}=D^2_\alpha \bar W^{12}=\bar D_{3\dot\alpha}
\bar W^{12}=0\,,\nn\\&&
D^1_\alpha W_{23}=\bar D_{2\dot\alpha}W_{23}=\bar
D_{3\dot\alpha}W_{23}=0\,;
\label{anal}
\eea
\item[(ii)] First-order harmonic shortness constraints,
\bea
&&
D^2_1\bar W^{12} =
D^1_2 \bar W^{12}=D^2_3 \bar W^{12}=D^1_3 \bar W^{12}=0\,,\nn\\&&
D^1_2 W_{23}=D^2_3 W_{23}=D^1_3 W_{23}=D^3_2 W_{23}=0\,;
\label{harm-short}
\eea
\item[(iii)] Second-order Grassmann linearity constraints,
\bea
&&(D^3)^2 \bar W^{12}=(\bar D_1)^2 \bar W^{12}=(\bar D_2)^2
\bar W^{12}=(\bar D_1 \bar D_2)\bar W^{12}=0\,,\nn\\&&
(D^2)^2W_{23}=(D^3)^2 W_{23}=(D^2 D^3)W_{23}=(\bar D_1)^2
W_{23}=0\,.
\label{linearity}
\eea
\end{itemize}
Altogether, the constraints (\ref{anal}), (\ref{harm-short}) and
(\ref{linearity}) kill all non-physical (auxiliary) field
components in $\bar W^{12}$ and $W_{23}$ and put the physical ones on shell\footnote{Besides eqs. (\ref{harm-short}),
the original constraints (\ref{constr}) imply some other relations of the first order in spinor derivatives, connecting $\bar W^{12}$ and $W_{23}$
with the remaining harmonic projections of $W_{kl}$ and $\bar W^{kl}$. These extra constraints can be used to deduce
the second-order constraints (\ref{linearity}) which, together with (\ref{anal}) and (\ref{harm-short}), form a closed set
of the harmonic superspace constraints on $\bar W^{12}$ and $W_{23}$ \cite{BS08}.}.
As a result, the superfield strengths
$\bar W^{12}$ and $W_{23}$ have the following component structure in the analytic
coordinates (\ref{anal-coord}),
\bea
\label{W-comp}
W_{23}&=&\varphi^1
+2i\theta_2^\alpha\bar\theta^{2\dot\alpha}\partial_{\alpha\dot\alpha}\varphi^1
-4i\theta_2^\alpha\bar\theta^{1\dot\alpha}\partial_{\alpha\dot\alpha}\varphi^2
-4i\theta_3^\alpha\bar\theta^{1\dot\alpha}\partial_{\alpha\dot\alpha}\varphi^3
\nn\\&&
+\,4i\theta_2^\alpha\theta_3^\beta F_{\alpha\beta}
+\bar\theta^{1\dot\alpha}\bar\lambda_{\dot\alpha}
+\theta_2^\alpha\lambda_{3\alpha}-\theta_3^\alpha\lambda_{2\alpha}
\nn\\&&
+\,2i\theta_2^\alpha\bar\theta^{2\dot\alpha}\bar\theta^{1\dot\beta}
\partial_{\alpha\dot\alpha}\bar\lambda_{\dot\beta}
+2i\theta_2^\beta\theta_3^\alpha \bar\theta^{2\dot\alpha}
\partial_{\alpha\dot\alpha}\lambda_{2\beta}
+4i\theta_2^\beta\theta_3^\alpha\bar\theta^{1\dot\alpha}
\partial_{\alpha\dot\alpha}\lambda_{1\beta}
\nn\\&&
+\,8\theta_2^\alpha\theta_3^\beta\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}
\partial_{\alpha\dot\alpha}\partial_{\beta\dot\beta}\varphi^3\,,
\nn\\
\bar W^{12}&=&\bar\varphi_3
-2i\theta_2^\alpha\bar\theta^{2\dot\alpha}\partial_{\alpha\dot\alpha}\bar\varphi_3
+4i\theta_3^\alpha\bar\theta^{1\dot\alpha}\partial_{\alpha\dot\alpha}\bar\varphi_1
+4i\theta_3^\alpha\bar\theta^{2\dot\alpha}\partial_{\alpha\dot\alpha}\bar\varphi_2
\nn\\&&
+\,4i\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}\bar F_{\dot\alpha\dot\beta}
+\theta_3^\alpha\lambda_\alpha
-\bar\theta^{2\dot\alpha}\bar\lambda^1_{\dot\alpha}
+\bar\theta^{1\dot\alpha}\bar\lambda^2_{\dot\alpha}
\nn\\&&
+\,2i\theta_2^\alpha\theta_3^\beta\bar\theta^{2\dot\alpha}
\partial_{\alpha\dot\alpha}\lambda_\beta
+2i\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}\theta_2^\alpha
\partial_{\alpha\dot\alpha}\bar\lambda^2_{\dot\beta}
+4i\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}\theta_3^\alpha
\partial_{\alpha\dot\alpha}\bar\lambda^3_{\dot\beta}
\nn\\&&
+\,8\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}
\theta_2^\alpha\theta_3^\beta
\partial_{\alpha\dot\alpha}\partial_{\beta\dot\beta}\bar\varphi_1
\,.
\label{W}
\eea
Here
\be
\varphi^I=u^I_i \varphi^i\,,\qquad
\bar\varphi_I=\bar u_I^i \bar\varphi_i\,,
\ee
and $\varphi^i$ is a triplet of physical scalars, $\square\varphi^i=0$.
The four spinor fields are comprised by the SU(3) singlet
$\lambda_\alpha$ and the triplet $\lambda_{I\alpha}=\bar
u_I^i\lambda_{i\alpha}$ which obey free equations of motion,
$\partial^{\alpha\dot\alpha}\lambda_\alpha=\partial^{\alpha\dot\alpha}\lambda_{i\alpha}=0$.
The fields $F_{\alpha\beta}=F_{(\alpha\beta)}$ and
$\bar F_{\dot\alpha\dot\beta}=\bar F_{(\dot\alpha\dot\beta)}$ are
spinorial components of the Maxwell field strength $F_{mn}=\partial_m A_n-\partial_n
A_m$, $\partial^m F_{mn}=0$.
The crucial feature of the $\cN=3$ harmonic superspace approach
is that one can relax some of the constraints (\ref{anal}),
(\ref{harm-short}), (\ref{linearity}) and express the superfield
strengths in terms of unconstrained off-shell gauge superfield
potentials \cite{IZ}. Consider the analytic superfields $V^1_2$
and $V^2_3$, $D^1_\alpha(V^1_2,\;V^2_3)=\bar
D_{3\dot\alpha}(V^1_2,\;V^2_3)=0\,$. They possess the following gauge
transformations
\be
\delta V^1_2=iD^1_2 \lambda\,,\qquad
\delta V^2_3=iD^2_3 \lambda\,,
\ee
with $\lambda$ being an analytic gauge superfield parameter.
Using these superfields, one constructs the non-analytic gauge
potentials $V^2_1$ and $V^3_2$ as solutions of the
zero-curvature equations \cite{IZ},
\be
D^1_2 V^2_1=D^2_1 V^1_2\,,\qquad
D^2_3 V^3_2=D^3_2 V^2_3\,.
\label{zero}
\ee
Finally, the gauge-invariant superfield strengths $\bar W^{12}$ and $W_{23}$ can be
expressed in terms of $V^2_1$ and $V^3_2$ as
\be
\bar W^{12}=-\frac14 D^{1\alpha}D^1_\alpha V^2_1\,,\qquad
W_{23}=\frac14 \bar D_{3\dot\alpha}\bar D_3^{\dot\alpha}V^3_2\,.
\label{W-anal}
\ee
It should be pointed out that the analyticity constraints
(\ref{anal}) are valid off shell while the other constraints
(\ref{harm-short}) and (\ref{linearity}) put the superfield
strengths on shell, except for the equations
$D^1_2 \bar W^{12}=0$ and $D^2_3 W_{23}=0$ which are also satisfied
off shell.
\subsection{Superconformal transformations in $\cN=3$ HSS}
The $\cN=3$ superconformal group SU(2,2$|$3), besides the $\cN=3$ super Poincar\'e transformations, contains
dilatation (with the parameter $a$), $\gamma_5$-transformation (with the parameter $b$),
conformal boosts (with the parameters $k_{\alpha\dot\alpha}$), conformal
supersymmetry (with the parameters $\eta^i_\alpha$, $\bar\eta_{i\dot\beta}$) and SU(3)
R-symmetry transformations (with the parameters $\lambda_i^j$, $\overline{\lambda_i^j}=-\lambda_j^i$,
$\lambda_i^i=0$). The realization of this supergroup on the analytic coordinates
(\ref{anal}) was found in \cite{GIO-N3},
\bea
\delta_{\rm sc} x_A^{\alpha\dot\alpha}&=&ax_A^{\alpha\dot\alpha}+
k_{\beta\dot\beta}x_A^{\alpha\dot\beta}x_A^{\beta\dot\alpha}-
4k_{\beta\dot\beta}\theta_2^\beta\bar\theta^{2\dot\alpha}
\theta_2^\alpha\bar\theta^{2\dot\beta}+
4ix_A^{\alpha\dot\beta}\bar\theta^{1\dot\alpha}\bar
u_1^i\bar\eta_{i\dot\beta}\nonumber\\
&& +\,2ix_{A-}^{\alpha\dot\beta}\bar\theta^{2\dot\alpha}
\bar u_2^i \bar\eta_{i\dot\beta}+
4ix_A^{\beta\dot\alpha}\theta_3^\alpha u^3_i\eta^i_\beta+
2ix_{A+}^{\beta\dot\alpha}\theta_2^\alpha u^2_i\eta^i_\beta\nonumber\\
&& -\,4i\lambda_i^j\theta_3^\alpha\bar\theta^{1\dot\alpha}u^3_j\bar u_1^i-
2i\lambda_i^j\theta_2^\alpha\bar\theta^{1\dot\alpha}u^2_j\bar u_1^i-
2i\lambda_i^j\theta_3^\alpha\bar\theta^{2\dot\alpha}u^3_j\bar
u_2^i\,,
\nn\\
\delta_{\rm sc}\theta_2^\alpha&=&(a/2+ib)\theta_2^\alpha+
k_{\beta\dot\beta}x_{A+}^{\alpha\dot\beta}\theta_2^\beta-
4i(\theta_2^\alpha u^2_i+\theta_3^\alpha u^3_i)\theta_2^\beta\eta^i_\beta
\nonumber\\
&& +\,x_{A+}^{\alpha\dot\beta}\bar u_2^i\bar\eta_{\dot\beta i}+
\lambda_i^j(\theta_2^\alpha u^2_j+\theta_3^\alpha u^3_j)\bar
u_2^i\,,
\nn\\
\delta_{\rm sc}\theta_3^\alpha&=&(a/2+ib)\theta_3^\alpha+
k_{\beta\dot\beta}x_{A-}^{\alpha\dot\beta}\theta_3^\beta-
4i\theta_3^\alpha\theta_3^\beta u^3_i\eta^i_\beta
+x_{A-}^{\alpha\dot\beta}\bar u_3^i\bar\eta_{\dot\beta i}+
\lambda_i^j\theta_3^\alpha u^3_j\bar u_3^i\,,
\nn\\
\delta_{\rm sc}\bar\theta^{1\dot\alpha}&=&(a/2-ib)\bar\theta^{1\dot\alpha}+
k_{\beta\dot\beta}x_{A+}^{\beta\dot\alpha}\bar\theta^{1\dot\beta}+
4i\bar\theta^{1\dot\beta}\bar\theta^{1\dot\alpha}\bar
u_1^i\bar\eta_{\dot\beta i}+x_{A+}^{\beta\dot\alpha}u^1_i\eta^i_\beta-
\lambda_i^j\bar\theta^{1\dot\alpha}\bar u_1^i u^1_j\,,
\nn \\
\delta_{\rm sc}\bar\theta^{2\dot\alpha}&=&(a/2-ib)\bar\theta^{2\dot\alpha}+
k_{\beta\dot\beta}x_{A-}^{\beta\dot\alpha}\bar\theta^{2\dot\beta}+
4i\bar\theta^{2\dot\beta}(\bar\theta^{1\dot\alpha}\bar u_1^i+
\bar\theta^{2\dot\alpha}\bar u_2^i)\bar\eta_{\dot\beta i}
\nonumber\\
&& +\,x_{A-}^{\beta\dot\alpha}u^2_i\eta^i_\beta-
\lambda_i^j(\bar\theta^{1\dot\alpha}\bar u_1^i +
\bar\theta^{2\dot\alpha}\bar u_2^i)u^2_j\,,
\label{dxA}
\end{eqnarray}
where $x_{A\pm}^{\alpha\dot\alpha}=x_{A}^{\alpha\dot\alpha}\pm
2i\theta_2^\alpha\bar\theta^{2\dot\alpha}$.
For preserving the $\cN=3$ harmonic analyticity, the harmonic variables should transform
according to the rules,
\be
\begin{array}{ll}
\delta_{\rm sc} u^1_i=u^2_i\lambda^1_2+u^3_i\lambda^1_3\,,\quad&
\delta_{\rm sc}\bar u_1^i=0\,,\\
\delta_{\rm sc} u_i^2=u_i^3\lambda^2_3\,,&
\delta_{\rm sc}\bar u_2^i=-\bar u^i_1\lambda^1_2\,,\\
\delta_{\rm sc} u^3_i=0\,,&
\delta_{\rm sc}\bar u_3^i=-\bar u_2^i\lambda^2_3-\bar u_1^i\lambda^1_3\,,
\end{array}
\label{du}
\ee
where
\be
\lambda^I_J=-4ik_{\beta\dot\beta}\theta_J^\beta\bar\theta^{I\dot\beta}-
4i(\bar\eta_{\dot\beta i}\bar\theta^{I\dot\beta}\bar u_J^i+
\theta_J^\beta\eta^i_\beta u^I_i)+u^I_i\bar u_J^j\lambda^i_j\,.
\label{lambda}
\ee
In this paper we will use the so-called passive form of superconformal
transformations of superfields, when the variation is taken at
different points, e.g., $\delta_{\rm sc} W \simeq W'(x')-W(x)$. In such an approach,
not only the superfields but also their derivatives, as well as the superspace measures, should
be varied while computing the superconformal transformations of the
superfield actions.
It is known \cite{hss-book} that
the analytic measure $d\zeta(^{33}_{11})du$ is invariant under
(\ref{dxA}) and (\ref{du}),
\be
{\rm Ber}\,\frac{\partial(x_A',\theta',u')}{\partial(x_A,\theta,u)}=1\,.
\label{Ber}
\ee
Using the coordinate transformations (\ref{dxA}) and (\ref{du}),
it is straightforward to find the superconformal variations of harmonic derivatives:
\be
\begin{array}{ll}
\delta_{\rm sc} D^1_2=-\lambda^1_2 S_1\,, & \delta_{\rm sc} D^2_1=(\lambda^1_1-\lambda^2_2)D^2_1\,,\\
\delta_{\rm sc} D^2_3=-\lambda^2_3 S_2\,, & \delta_{\rm sc} D^3_2=(\lambda^2_2-\lambda^3_3)D^3_2\,,\\
\delta_{\rm sc} D^1_3=\lambda^1_2D^2_3-\lambda^2_3D^1_2-\lambda^1_3(S_1+S_2)\,,\quad &
\delta_{\rm sc} D^3_1=(\lambda^1_1-\lambda^3_3)D^3_1+\lambda^2_1D^3_2-\lambda^3_2D^2_1\,,\\
\delta_{\rm sc} D^1_1=\delta_{\rm sc} D^2_2=\delta_{\rm sc} D^3_3=0\,,&
\delta_{\rm sc} S_1=\delta_{\rm sc} S_2=0\,.
\end{array}
\label{dD}
\ee
Recall that the gauge covariant harmonic derivatives involve the gauge superfield prepotentials
\be
\nabla^I_J=D^I_J+iV^I_J\,.
\label{cov-gauge}
\ee
Requiring the lengthened derivatives (\ref{cov-gauge}) to be superconformally covariant, with taking into account the transformations (\ref{dD}),
implies the following transformation laws for the gauge prepotentials:
\be
\begin{array}{ll}
\delta_{\rm sc} V^1_2=0\,, & \delta_{\rm sc} V^2_1=(\lambda^1_1-\lambda^2_2)V^2_1\,,\\
\delta_{\rm sc} V^2_3=0\,, & \delta_{\rm sc} V^3_2=(\lambda^2_2-\lambda^3_3)V^3_2\,,\\
\delta_{\rm sc} V^1_3=\lambda^1_2 V^2_3-\lambda^2_3 V^1_2\,, \quad&
\delta_{\rm sc} V^3_1=(\lambda^1_1-\lambda^3_3)V^3_1+
\lambda^2_1V^3_2-\lambda^3_2V^2_1\,.
\end{array}
\label{dV}
\ee
Note that the superconformal variations of the analytic gauge
superfields $V^1_2$, $V^2_3$ and $V^1_3$ were earlier given
in \cite{GIO-N3,hss-book}, while the transformations of the
non-analytic gauge superfields $V^2_1$, $V^3_2$ and $V^3_1$ were not presented before.
Using (\ref{dxA}) and (\ref{du}) it is also easy to find the
superconformal transformations of the covariant spinor derivatives
(\ref{anal-deriv}),
\bea
\delta_{\rm sc} D^1_\alpha&=&(-a/2-ib-\lambda^1_1)D^1_\alpha+B_\alpha^\beta D^1_\beta\,,
\nn\\
\delta_{\rm sc}\bar D_{3\dot\alpha}&=&(-a/2+ib+\lambda^3_3)\bar D_{3\dot\alpha}+
\bar B_{\dot\alpha}^{\dot\beta}\bar D_{3\dot\beta}\,,
\label{delta-D}
\eea
where $\lambda^1_1$ and $\lambda^3_3$ were defined in
(\ref{lambda}) and
\bea
B_\alpha^\beta&=&-k_{\alpha\dot\beta}(x_{A+}^{\beta\dot\beta}+
4i\theta_1^\beta\bar\theta^{1\dot\beta})-4i\theta_I^\beta
u^I_j\eta^j_\alpha\,,\nn\\
\bar B_{\dot\alpha}^{\dot\beta}&=&-k_{\beta\dot\alpha}
(x_{A-}^{\beta\dot\beta}-4i\theta_3^\beta\bar\theta^{3\dot\beta})-
4i\bar\theta^{I\dot\beta}\bar u_I^j\bar\eta_{\dot\alpha j}\,.
\eea
It is worth pointing out that the spinor derivatives $ D^1_\alpha $ and $\bar D_{3\dot\alpha}$ are not mixed under the superconformal
transformations.
Finally, using the variations of the gauge prepotentials
(\ref{dV}) and derivatives (\ref{delta-D}), we can find the
superconformal transformations of the superfield strengths
(\ref{W-anal}),
\be
\delta_{\rm sc} W_{23}=A W_{23}\,,\qquad
\delta_{\rm sc} \bar W^{12}=\bar A \bar W^{12}\,,
\label{dW}
\ee
where
\be
A=-a+2ib+\lambda^2_2+\lambda^3_3+\bar
B^{\dot\alpha}_{\dot\alpha}\,,\qquad
\bar A=-a-2ib-\lambda^1_1-\lambda^2_2+B^\alpha_\alpha\,.
\label{A}
\ee
One can check that the superfields $A$ and $\bar A$ are analytic,
\be
D^1_\alpha (A,\;\bar A)=\bar D_{3\dot\alpha}(A,\;\bar A)=0\,.
\ee
Hence, the transformations (\ref{dW}) preserve analyticity.
\section{Superconformal effective action}
\subsection{Non-superconformal $F^4$ term}
The $\cN=3$ supersymmetric completion of the fourth-order term
in the Born-Infeld action was constructed in \cite{IZ},
\be
S_4=\frac1{32}\int d\zeta(^{33}_{11})du\,
\frac{(\bar W^{12} W_{23})^2}{(\bar\Lambda\Lambda)^2}\,.
\label{S4}
\ee
Here $\Lambda$ is a coupling constant of dimension one in mass units, which is
introduced to ensure the correct dimension of the integrand.
The analytic measure is defined as follows \cite{IZ,BISZ},
\be
d\zeta(^{33}_{11})=\frac1{16^2}d^4x_A(D^3)^2(D^2)^2(\bar D_1)^2(\bar
D_2)^2\,.
\label{measure}
\ee
The analytic measure is dimensionless,
$[d\zeta(^{33}_{11})du]=0\,$, and $[\bar W^{12}]=[W_{23}]=1\,$.
With this normalization of the analytic measure, it is
straightforward to check that, along with other component terms, the
action (\ref{S4}) yields the standard $F^4$ term,
\be
S_4=\frac12\int d^4x\,\frac{F^2\bar
F^2}{(\bar\Lambda\Lambda)^2}+\ldots\,.
\label{F4}
\ee
Consider now the superconformal variation of the action (\ref{S4}),
\be
\delta_{\rm sc} S_{4}=\frac1{16}\int d\zeta(^{33}_{11})du(A+\bar A)
\frac{(\bar W^{12} W_{23})^2}{(\bar\Lambda\Lambda)^2}\,,
\label{dS4}
\ee
where we have used the variations of the superfield strengths
(\ref{dW}) and the invariance of the analytic measure (\ref{Ber}).
Here $A$ and $\bar A$ are superfields (\ref{A}) collecting
the constant parameters of the superconformal transformations
(\ref{dxA}) and (\ref{du}). The variation (\ref{dS4}) is non-zero, hence the action (\ref{S4})
is not superconformal.
\subsection{Scale and $\gamma_5$ invariant $F^4/X^4$ term}
Our aim here is to find a superconformal generalization of the action (\ref{S4}).
In what follows we will denote this superconformal action by $\Gamma$ (to stress that
it is a part of the $\cN=3$ SYM low-energy effective action). The action $\Gamma$
should meet the following criteria:
\begin{enumerate}
\item It should be a local functional defined on the analytic superspace and constructed out of the superfield strengths
$\bar W^{12}$ and $W_{23}$ without derivatives on them,
\be
\Gamma=\int d\zeta(^{33}_{11}) du\,{\cal H}^{11}_{33}(\bar W^{12},W_{23})\,.
\label{3.5}
\ee
The analytic Lagrangian density ${\cal H}^{11}_{33}$ is an arbitrary function of its
arguments, such that its external harmonic U(1) charges cancel those of the analytic integration measure.
This is the most general form of the superspace action yielding terms with four-derivatives in components, since
the analytic measure (\ref{measure}) contains just eight spinor derivatives which can produce four space-time ones
on the component fields.
\item
The action $\Gamma$ should be invariant under the superconformal
transformations (\ref{dW}),
\be
\delta_{\rm sc} \Gamma=0\,.
\ee
As a weaker requirement, in this subsection we will employ only the scale- and
$\gamma_5$-transformations out of the full SU(2,2$|$3)
superconformal group. We will show that this is sufficient to uniquely specify the structure of the action. The check of the full
superconformal symmetry will be performed in the next subsection.
\item
In the component-field expansion the action $\Gamma$ should
reproduce the scale- and SU(3)-invariant
$F^4/X^4$ term (\ref{F4}),
\be
\int d^4x\frac{F^2\bar F^2}{(\varphi^i\bar\varphi_i)^2}\,.
\ee
\item
We are interested in the low-energy effective action for massless
fields, with massive ones being integrated out. The massive fields appear
in the Coulomb branch when the gauge symmetry is broken down
spontaneously. For instance, the SU(2) gauge symmetry is broken down to U(1)
when the scalar field corresponding to the Cartan subalgebra of
su(2) acquire non-trivial vevs,
\be
c^i=\langle \varphi^i \rangle\ne0\,,\qquad
\bar c_i=\langle \bar \varphi_i \rangle \ne0\,.
\label{c}
\ee
However, the effective action should be independent of any particular choice
of these constants,
\be
\Gamma({c'}^i,\bar c'_j)=\Gamma(c^i,\bar c_j)\,, \qquad c^i\bar c_i \ne0\,,
\ee
because such a dependence would break superconformal invariance of the action.
\item
Finally, we simplify the problem by considering only that part of
the action (\ref{3.5}) which does not vanish on the mass shell, i.e.,
we will assume that the superfield strengths obey the constraints
(\ref{anal})--(\ref{linearity}). We will neglect all terms in
the action $\Gamma$ which vanish when these constraints are imposed.
As a consequence, one is free to add to $\Gamma$ or to subtract from it
the following expressions which vanish on the mass shell,
\bea
\int d\zeta(^{33}_{11})\, \bar W^{12}{\cal F}(W_{23})&\propto&
\int d^4x (D^3)^2 (D^2)^2 (\bar D_1)^2[{\cal F}(W_{23}) (\bar D_2)^2 \bar
W^{12}] \simeq 0\,,\nn\\
\int d\zeta(^{33}_{11})\, W_{23}{\cal F}( \bar W^{12})&\propto&
\int d^4x (D^3)^2 (\bar D_2)^2 (\bar D_1)^2[{\cal F}( \bar W^{12})
(D^2)^2 W_{23}] \simeq 0\,.\nn\\
\label{prop}
\eea
Here ${\cal F}(W)$ is an arbitrary function of its argument.
We will frequently employ this property while deriving the action.
\end{enumerate}
Now we shall turn to constructing the action $\Gamma$ which obeys the requirements and properties listed above.
As the first step, we introduce the shifted scalar fields, $\phi^i$ and
$\bar\phi_i$,
\be
\varphi^i=c^i+\phi^i\,,\quad
\bar\varphi_i=\bar c_i+\bar \phi_i\,,\qquad
\langle \phi^i \rangle=\langle \bar\phi_i \rangle=0\,.
\label{vev-shift}
\ee
Next, we define the harmonic projections of these vev
constants
\be
c^1=u^1_i c^i\,,\quad c^2=u^2_i c^i\,\quad
c^3=u^3_i c^i\,,\qquad
\bar c_1 =\bar u_1^i \bar c_i\,,\quad
\bar c_2 =\bar u_2^i \bar c_i\,,\quad
\bar c_3 = \bar u_3^i \bar c_i\,.
\label{cccc}
\ee
Using these objects, we introduce the shifted superfield
strengths, $\bar\omega^{12}$ and $\omega_{23}$,
\be
\bar W^{12}=\bar c_3+\bar\omega^{12}\,,\qquad
W_{23}=c^1+\omega_{23}\,.
\label{shift}
\ee
Under the scale and $\gamma_5$ transformations these shifted
superfields transform inhomogeneously,
\be
\delta_{\rm sc} \bar\omega^{12}=\bar A\bar c_3 +\bar A\bar\omega^{12}\,,\qquad
\delta_{\rm sc} \omega_{23}=Ac^1+A\omega_{23}\,,
\label{scale}
\ee
where $A=-a+2ib$. The case of generic $A$ and $\bar A$ defined in (\ref{A}) will be considered in the next
subsection.
We point out that on shell, when the relations (\ref{prop}) are valid, the non-superconformal action (\ref{S4}) can be
rewritten in terms of $\bar \omega^{12}$ and $\omega_{23}$ as
\be
S_4=\frac1{32}\int d\zeta(^{33}_{11})du\,
\frac{(\bar \omega^{12} \omega_{23})^2}{(c^i\bar c_i)^2}\,.
\label{S4_}
\ee
Here we substituted $(c^i\bar c_i)^2$ in the denominator instead of
$(\bar\Lambda\Lambda)^2$, because no other dimensionful constants besides the vevs $c^i$ can be
present in the superconformal case.
We search for a superconformal generalization of the action (\ref{S4_}) in the form
\be
\Gamma=\frac\alpha{8}\int d\zeta(^{33}_{11})du\frac{(\bar \omega^{12}
\omega_{23})^2}{(c^i\bar c_i)^2}
H\left(\frac{\bar \omega^{12}c^3}{c^i \bar c_i},
\frac{\omega_{23}\bar c_1}{c^i\bar c_i}\right),
\label{G}
\ee
where $H(x,y)$ is some function to be determined and $\alpha$ is a
dimensionless coupling constat.
The arguments $\frac{\bar \omega^{12}c^3}{c^i \bar
c_i}$ and $\frac{\omega_{23}\bar c_1}{c^i\bar c_i}$ of the function $H$ are chargeless and dimensionless.
We assume that the function $H$ has a regular power expansion with respect to its arguments,
\be
H(x,y)=\sum_{m,n=0}^\infty
\alpha_{m,n} x^m y^n\,,
\label{H}
\ee
with undefined coefficients $\alpha_{m,n}$.
The reality of the action (\ref{G}) under complex conjugation
implies the symmetry of this function, $H(x,y)=H(y,x)\,$, whence $\alpha_{m,n}=\alpha_{n,m}\,$.
Reordering the summation in (\ref{H}), it is convenient to
represent (\ref{G}) as
\be
\Gamma=\sum_{n=0}^\infty\Gamma_n\,,\qquad
\Gamma_n=\frac\alpha{8}\int d\zeta(^{33}_{11})du
\frac{(\bar \omega^{12}\omega_{23})^2}{(c^i\bar c_i)^2}
\sum_{i=0}^{n}
\alpha_{i,n-i}
\left(\frac{\bar\omega^{12}c^3}{c^i\bar c_i}\right)^i
\left(\frac{\omega_{23}\bar c_1}{c^i\bar c_i}\right)^{n-i}.
\label{Gn}
\ee
The invariance of the action (\ref{Gn}) under the transformations
(\ref{scale}) can be ensured order by order, i.e., the
non-vanishing terms from $\delta_{\rm sc} \Gamma_{n}$ are required to be cancelled by similar terms from $\delta_{\rm sc} \Gamma_{n+1}$,
and so forth. This recurrence procedure imposes severe restrictions on the coefficients
$\alpha_{m,n}$. The technical details of this procedure are given in the Appendix A, with the following result:
\be
\alpha_{m,n}=(-1)^{m+n}\frac{(m+n+2)!}{(n+2)n!(m+2)m!}\,.
\label{alpha}
\ee
With these coefficients, the series (\ref{H}) can be summed up as follows,
\be
H(x,y)=
\frac{\ln (1 + x + y)}{x^2y^2}+
\frac{1}
{xy( 1 + x +
y ) } -
\frac{\ln (1 + x)}{x^2 y^2} -
\frac{\ln (1 + y)}{x^2y^2} \,.
\label{HH}
\ee
We point out that this function is regular at the origin,
\be
\lim_{x,y\to0}H(x,y)=\frac12\,.
\ee
Hence, the action (\ref{G}) with this function is well-defined and
the harmonic integral does not encounter any singularities.
The contributions from the last two terms in (\ref{HH}) to the
action (\ref{G}) vanish on shell due to the properties
(\ref{prop})\footnote{The properties (\ref{prop}) are valid essentially on shell.
Therefore the last two terms in (\ref{HH}) can be neglected only
on shell although they can be important for the off-shell
completion of the action.}. Therefore, the on-shell effective
action can be rewritten in the following explicit form
\be
\Gamma=\frac\alpha{8}\int d\zeta(^{33}_{11})du\left[\frac{(c^i\bar c_i)^2}{c^3c^3 \bar c_1 \bar c_1}
\ln\left(1+\frac{\bar\omega^{12}c^3}{c^i\bar c_i}
+\frac{\omega_{23}\bar c_1}{c^i\bar c_i}\right)
+\frac{(c^i\bar c_i)\bar\omega^{12}\omega_{23}}{c^3\bar c_1(c^i\bar c_i
+\bar\omega^{12}c^3+\omega_{23}\bar c_1)}
\right].
\label{Gconf}
\ee
Although the charged objects $c^3$ and $\bar c_1$ appear in the
denominators, they do not lead to the divergent harmonic
integrals. It can be explicitly checked that upon passing
to the component form of the action (\ref{Gconf}), all dangerous
terms with divergent harmonic integrals vanish after performing the integration
over the Grassmann variables. Some component terms of this action
will be studied in the next Section.
\subsection{Complete $\cN=3$ superconformal symmetry}
Now we consider the transformations (\ref{dW})
which include all parameters of the superconformal
transformations. The corresponding variations (\ref{scale}) of the
shifted superfield strengths $\bar \omega^{12}$ and $\omega_{23}$
read
\bea
\delta_{\rm sc} \bar \omega^{12}&=&A\bar\omega^{12}+A\bar c_3+\lambda^2_3
\bar c_2+\lambda^1_3 \bar c_1\,,\nn\\
\delta_{\rm sc} \omega_{23}&=&\bar A \omega_{23}+\bar A c^1
-\lambda^1_2 c^2-\lambda^1_3 c^3\,,
\eea
where $A$ and $\bar A$ are given in (\ref{A}) and $\lambda^I_J$
are defined in (\ref{lambda}). Under these transformations the
action (\ref{G}) varies as
\bea
\delta_{\rm sc} \Gamma&=&\frac\alpha{8}\int d\zeta(^{33}_{11})du\, (\bar\omega^{12}\omega_{23})^2
[\frac{2}x H(x,y)+H'_x(x,y)][Ax+Ac^3\bar c_3+\lambda^2_3 c^3\bar c_2
+\lambda^1_3 c^3 \bar c_1]
\nn\\
&&+\,\frac\alpha{8}\int d\zeta(^{33}_{11})du\, (\bar\omega^{12}\omega_{23})^2
[\frac{2}y H(x,y)+H'_y(x,y)][\bar Ay+\bar Ac^1\bar c_1-\lambda^1_2 c^2\bar
c_1 -\lambda^1_3 c^3 \bar c_1]\,.\nn\\
\label{var1}
\eea
For simplicity we set here $c^i\bar c_i=1\,$, so $x=\bar\omega^{12}c^3$, $y=\omega_{23}\bar c_1$. The
first and second lines in (\ref{var1}) are complex-conjugated to each other.
Given the explicit form (\ref{HH}) of the function $H(x,y)$, it is
easy to check that it solves the following differential equations
\bea
\frac2xH(x,y)+H'_x(x,y)&=&\frac{1}{x(1+x)(1+x+y)^2}\,,\nn\\
\frac2yH(x,y)+H'_y(x,y)&=&\frac{1}{y(1+y)(1+x+y)^2}\,.
\label{relH}
\eea
Taking them into account, we are going to show that the integrand in (\ref{var1}) is a total harmonic derivative,
so the variation (\ref{var1}) vanishes.
To this end, we introduce the auxiliary functions $f(x,y)$ and $\tilde{f}(x,y)$:
\bea
f(x,y)&=&\frac{1}{y(y+1)(x+y+1)}+\frac{\ln(1+x+y)}{xy^2}
-\frac{\ln(1+x)}{xy^2}-\frac{\ln(1+y)}{xy^2}\,,\\
\tilde f(x,y)&=&f(y,x)=
\frac{1}{x(x+1)(x+y+1)}+\frac{\ln(1+x+y)}{yx^2}
-\frac{\ln(1+y)}{yx^2}-\frac{\ln(1+x)}{yx^2}\,.\nn
\label{f}
\eea
They possess the following properties
\bea
xf'_x+f&=&-\frac1{(1+x)(1+x+y)^2}=-(xH'_x+2H)\,,
\label{prop0.1}\\
xf'_x+yf'_y+3f&=&\frac1{x(1+x)(1+x+y)^2}-\frac1{x(1+y)^2}
=(H'_x+\frac 2xH)+\ldots\,,\label{prop1.1}\\
y\tilde f'_y+\tilde f&=&-\frac1{(1+y)(1+x+y)^2}=-(yH'_y+2H)\,,\\
y\tilde f'_y+x\tilde f'_x +3\tilde f&=&
\frac1{y(1+y)(1+x+y)^2}-\frac1{y(1+x)^2}=(H'_y+\frac 2x
H)+\ldots\,.
\label{prop1}
\eea
Here dots stand for the terms integrals of which over the
analytic superspace with the weight
$(\bar\omega^{12}\omega_{23})^2$ are on-shell vanishing due to the relations (\ref{prop}).
Up to these terms, the
equations (\ref{prop0.1})--(\ref{prop1}) allow one to deduce the following relations
\bea
-D^2_3 (f(x,y)c^3 \bar c_2 A)- D^1_3(f(x,y)c^3 \bar c_1 A)
&=&(H'_x+\frac 2x H)(Ax+Ac^3\bar c_3)
\nn\\&&
-\,f(x,y)c^3 \bar c_2
\lambda^2_3
-f(x,y)c^3\bar c_1 \lambda^1_3\,,\nn\\
D^1_2(\tilde f(x,y)c^2\bar c_1\bar A)
+D^1_3(\tilde f(x,y)c^3\bar c_1\bar A)
&=&(H'_y+\frac2yH)(\bar A y+\bar A c^1\bar c_1)
\nn\\&&
+\,\tilde f(x,y)c^2\bar c_1\lambda^1_2
+\tilde f(x,y)c^3\bar c_1\lambda^1_3\,.
\label{rel1}
\eea
Here we made use of the following simple identities
\be
\lambda^1_2=D^1_2\bar A\,,\quad
\lambda^2_3=D^2_3A\,,\quad
\lambda^1_3=D^1_3A=D^1_3\bar A\,,
\ee
as well as of the convention $c^i\bar c_i=1\,$.
Next, we introduce the functions
\bea
g(x,y)&=&\frac1{y(1+y)^2(1+x+y)}-\frac1{y(x+1)}\,,\\
\tilde g(x,y)&=&g(y,x)=\frac1{x(1+x)^2(1+x+y)}-\frac1{x(y+1)}\,,
\label{g}
\eea
with the properties
\bea
xg'_x+g&=&\frac1{y(1+y)(1+x+y)^2}-\frac1{y(1+x)^2}
=H'_y+\frac2yH+\ldots\,,\\
y\tilde g'_y+\tilde
g&=&\frac1{x(1+x)(1+x+y)^2}-\frac1{x(1+y)^2}
=H'_x+\frac 2xH+\ldots\,,\\
g(x,y)-\tilde g(x,y)&=&(H'_x+\frac 2x H)-(H'_y+\frac 2y H)\,.
\eea
Here, as in (\ref{prop1.1}) and in (\ref{prop1}), the dots stand for the terms vanishing on shell after integration over the
analytic superspace with the weight $(\bar
\omega^{12}\omega_{23})^2$. Up to these terms, we obtain the
following relation
\bea
-D^1_2(\lambda^2_3\tilde g(x,y)c^3\bar c_1)-D^2_3(\lambda^1_2 g(x,y)c^3\bar c_1)
&=&(H'_x+\frac2xH)\lambda^2_3 c^3\bar c_2
-(H'_y+\frac 2yH)\lambda^1_2c^2\bar c_1
\nn\\&&+\,[(H'_x+\frac2xH)-(H'_y+\frac 2yH)]
\lambda^1_3 c^3\bar c_1\,.
\nn\\
\label{rel2}
\eea
Finally, introduce the functions
\bea
h(x,y)&=&-\frac1{(1+x)y}+\frac{\ln(1+x)}{xy^2}+\frac{\ln(1+y)}{xy^2}
-\frac{\ln(1+x+y)}{xy^2}\,,
\\
\tilde h(x,y)&=&h(y,x)=
-\frac1{(1+y)x}+\frac{\ln(1+y)}{yx^2}+\frac{\ln(1+x)}{yx^2}
-\frac{\ln(1+x+y)}{yx^2}\,,
\label{h}
\eea
with the properties
\bea
&h(x,y)+yh'_y(x,y)=f(x,y)\,,\qquad
\tilde h(x,y)+x\tilde h'_x(x,y)=\tilde f(x,y)\,,&\\
&h-\tilde h=\tilde f-f\,.&
\eea
These properties allow one to derive the following relation
\be
-D^1_2(\lambda^2_3 h(x,y)c^3\bar c_1)-D^2_3(\lambda^1_2 \tilde h(x,y)c^3\bar c_1)
=f\lambda^2_3 c^3\bar c_2
-\tilde f\lambda^1_2c^2\bar c_1
+(f-\tilde f)\lambda^1_3 c^3\bar c_1\,.
\label{rel3}
\ee
Now we put together the relations (\ref{rel1}), (\ref{rel2}) and
(\ref{rel3}) and observe that the variation (\ref{var1}) can
be represented as a linear combination of harmonic
derivatives acting on the quantities which are expressed through the
functions (\ref{f}), (\ref{g}) and (\ref{h}),
\bea
\delta_{\rm sc} \Gamma&=&\frac\alpha{8}\int d\zeta(^{33}_{11})du\, (\bar\omega^{12}\omega_{23})^2
\bigg\{
D^1_2(\tilde f c^2 \bar c_1 \bar A)
- D^2_3( f c^3 \bar c_2 A)
+D^1_3(\tilde f c^3 \bar c_1\bar A- f c^3 \bar c_1 A)
\nn\\&&
-\,D^1_2[(\tilde g+h)\lambda^2_3 c^3\bar c_1]-D^2_3[(g+\tilde h)\lambda^1_2 c^3\bar c_1]
\bigg\}.\label{344}
\eea
The variation (\ref{344}) vanishes as an integral of total
harmonic derivative. This proves the invariance of the action
(\ref{Gconf}) under the full SU(2,2$|$3) superconformal group\footnote{Note that (\ref{Gconf}) is SU(2,2$|$3) invariant for any $c^i\neq 0\,$,
without any restriction on the norm $c^i\bar c_i$ which was put equal to $1$ in the above consideration merely for convenience.}.
\subsection{Independence of the choice of vacua}
By construction, the effective action (\ref{G}) with the function
(\ref{HH}) is meaningful only on the Coulomb branch of the $\cN=3$
SYM theory. This is manifested in the explicit presence of non-zero
vev constants $c^i$ and $\bar c_i$ in the Lagrangian in (\ref{G}). However, the
action itself should be independent of any particular choice of these
constants, except for the point $c^i=0$ at which the effective action
is singular.
Let us rewrite (\ref{G}) in terms of the original
(non-shifted) superfield strengths $\bar W^{12}$ and $W_{23}$
\be
\Gamma[\bar W^{12},W_{23};c^i,\bar c_i]=\frac\alpha{8}\int d\zeta(^{33}_{11})du\frac{(\bar W^{12}-\bar c_3)^2
(W_{23}-c^1)^2}{(c^i\bar c_i)^2}
H\left(c^3\frac{\bar W^{12}-\bar c_3}{c^i \bar c_i},
\bar c_1\frac{W_{23}-c^1}{c^i\bar c_i}\right).
\label{GG}
\ee
In the previous subsection we proved that this action is invariant
under the full SU(2,2$|$3) superconformal group. Taking into account that the analytic integration measure is SU(2,2$|$3) invariant by itself,
the property of superconformal invariance of the action can be written in the finite form as
\be
\Gamma[\bar W^{12},W_{23};c^i,\bar c_i] = \Gamma'[\bar W^{12}{}',W_{23}{}';c^i,\bar c_i] = \Gamma[\bar W^{12}{}',W_{23}{}';c^i,\bar c_i]\,.
\ee
In particular,
consider scale and $\gamma_5$ transformations of the superfield
strength in the finite form,
\be
\bar W^{12}\to e^{\bar A} \bar W^{12}\,,\qquad
W_{23}\to e^A W_{23}\,,
\label{scale-finite}
\ee
where $A=-a+2ib$. The transformation of the action (\ref{GG})
under (\ref{scale-finite}) can be represented as
\bea
\Gamma[\bar W^{12},W_{23};c^i,\bar c_i] &=& \Gamma[e^{\bar A} \bar W^{12},e^A W_{23};c^i,\bar c_i] \nonumber \\
&=&\frac\alpha{8}
\int d\zeta(^{33}_{11})du\frac{(\bar W^{12}-e^{-\bar A}\bar c_3)^2
(W_{23}-e^{-A}c^1)^2}{(e^{-A-\bar A}c^i\bar c_i)^2}
\nn\\&&\times
\,H\left(e^{-A}c^3\frac{\bar W^{12}-e^{-\bar A}\bar c_3}{e^{-A-\bar A}c^i \bar c_i},
e^{-\bar A}\bar c_1\frac{W_{23}-e^{-A}c^1}{e^{-A-\bar A}c^i\bar c_i}\right).
\eea
So, all $A$-dependence is absorbed into the vev constants,
$c^i\to e^{-A}c^i$, $\bar c_i\to e^{-\bar A}\bar c_i$. Hence, the
superconformal invariance of the action (\ref{GG}) implies its
independence of complex rescalings of the vev constants,
\be
\Gamma[\bar W^{12},W_{23};c^i,\bar c_i]
=\Gamma[e^{\bar A} \bar W^{12},e^A W_{23};c^i,\bar c_i]
=\Gamma[ \bar W^{12},W_{23};e^{-\bar A}c^i,e^{-A}\bar c_i]\,.
\ee
In a similar way, one can prove that the action
(\ref{GG}) is independent of the parameters of finite SU(3) rotations of the vev constants,
\be
\Gamma[\bar W^{12},W_{23};c^i,\bar c_i]
=\Gamma[ \bar W^{12},W_{23};\Lambda^i_j c^j,\bar \Lambda_i^j\bar
c_j]\,,
\ee
where $\Lambda^i_j$ are SU(3) matrices. As a result, the action
(\ref{GG}) is independent of any particular choice of the vacuum
$c^i$, $c^i\ne0\,$.
Perhaps, it make sense to give a more detailed proof of the latter statement. It goes as follows. Let us assume, without loss of generality,
that $c^3\neq 0\,$. Then, using the coset SU(3)/[U(1)$\times $SU(2)] transformations with a constant SU(2) doublet as parameters, one can
cast $c^i$ in the form
$c^i = (0,0,c^3)$. The constant $c^3$ can be made real by making use of the residual U(1) transformation
(a combination of the $\gamma_5$ transformations
and those of U(1) from the denominator of SU(3)/[U(1)$\times $SU(2)]). Finally, it can be rescaled to any non-zero value,
keeping in mind the independence of the action of the rescalings of the vev constants.
\section{Component structure}
\subsection{$F^4/X^4$ term}
To derive this term from the effective action (\ref{G}), it is
sufficient to consider only constant Maxwell and scalar fields,
omitting all other components in (\ref{W-comp}),
\be
\hat{\bar\omega}^{12}=u^1_i \phi^i+4i\theta_2^\alpha\theta_3^\beta
F_{\alpha\beta}\,,\qquad
\hat\omega_{23}=\bar u_3^i
\bar\phi_i+4i\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}\bar
F_{\dot\alpha\dot\beta}\,.
\ee
Substituting these superfields into (\ref{G}), we integrate over the
Grassmann variables to obtain
\be
\Gamma_{F^4/X^4}=\frac\alpha{2}\int d^4x du\, F^2\bar F^2
\sum_{m,n=0}^\infty \frac{(m+1)(n+1)(m+n+2)!(-1)^{m+n}}{m!n!}
(\bar \phi_3 c^3)^m (\phi^1 \bar c_1)^n\,.
\label{G-comp}
\ee
Here we used the series expansion (\ref{H}) for the function
$H$ with the coefficients given by (\ref{alpha}). In this
subsection we assume $c^i\bar c_i=1$ for simplicity
and use the notation
$F^2=F^{\alpha\beta}F_{\alpha\beta}$\,,
$\bar F^2=\bar F^{\dot\alpha\dot\beta}\bar
F_{\dot\alpha\dot\beta}$\,.
It is convenient to represent (\ref{G-comp}) as a sum of two
terms,
\be
\Gamma_{F^4/X^4}=\frac\alpha2\int d^4x\, F^2\bar F^2(T_1+T_2)\,,
\ee
where
\bea
T_1&=&\int du\,
\sum_{n=0}^\infty \sum_{m=0}^n \frac{(m+1)(n+1)(m+n+2)!(-1)^{m+n}}{m!n!}
(\phi^1 \bar c_1)^n (\bar \phi_3 c^3)^m \,,\nn\\
T_2&=&\int du\,
\sum_{n=0}^\infty \sum_{m=n+1}^\infty \frac{(m+1)(n+1)(m+n+2)!(-1)^{m+n}}{m!n!}
(\phi^1 \bar c_1)^n(\bar \phi_3 c^3)^m \,.
\eea
The reason for this separation is that $m\leq n$ in $T_1$ while $m>n$
in $T_2$. Therefore, for each of these terms we can apply the
equation (\ref{hint}) for the harmonic integrals,
\bea
T_1&=&2\sum_{n=0}^\infty \sum_{m=0}^n \sum_{l=0}^m
\frac{(m+n-l+1)!(-1)^{m+n+l}}{l!(n-l)!(m-l)!}
(\phi^i\bar\phi_i)^l (\phi^i\bar
c_i)^{n-l}(c^i\bar\phi_i)^{m-l}\,,\nn\\
T_2&=&2\sum_{n=0}^\infty \sum_{m=n+1}^\infty \sum_{l=0}^n
\frac{(m+n-l+1)!(-1)^{m+n+l}}{l!(n-l)!(m-l)!}
(\phi^i\bar\phi_i)^l (\phi^i\bar
c_i)^{n-l}(c^i\bar\phi_i)^{m-l}\,.
\eea
Changing the order of summation, these
terms can be rewritten as
\bea
T_1&=&2\sum_{l,m=0}^\infty \sum_{n=m}^\infty
\frac{(n+m+l+1)!(-1)^{m+n+l}}{l!m!n!}(\phi^i\bar\phi_i)^l
(\phi^i\bar\phi_i)^n (c^i\bar\phi_i)^m\,,\nn\\
T_2&=&2\sum_{l,m=0}^\infty \sum_{n=0}^{m-1}
\frac{(n+m+l+1)!(-1)^{m+n+l}}{l!m!n!}(\phi^i\bar\phi_i)^l
(\phi^i\bar\phi_i)^n (c^i\bar\phi_i)^m\,.
\eea
Putting these two expressions together, we find
\bea
T_1+T_2&=&2\sum_{m,n,k=0}^\infty
\frac{(-1)^{m+n+k}(m+n+k+1)!}{m!n!k!}
(c^i\bar\phi_i)^{m}(\bar c_i\phi^i)^n(\bar\phi_i\phi^i)^{k}
\nn\\&
=&\frac2{(1+c^i\bar\phi_i+\bar c_i \phi^i+\phi^i\bar\phi_i)^2}
=\frac2{(\varphi^i\bar\varphi_i)^2}\,.
\eea
As a result, the $F^4/X^4$ term in the effective action reads
\be
\Gamma_{F^4/X^4}=\alpha\int d^4x\frac{F^2\bar
F^2}{(\varphi^i\bar\varphi_i)^2}\,. \label{F4X}
\ee
This expression is explicitly scale and U(3) invariant, as is
expected.
It is a highly non-trivial and remarkable phenomenon that the vev constants $c^i$ and the
shifted scalars $\phi^i$ have combined into the initial scalar
fields $\varphi^i$, (\ref{vev-shift}), after doing the Grassmann and harmonic integrals which
is a rather involved procedure in its own.
This confirms the independence of the action (\ref{G}) of any particular choice of
the vacua, the fact that was proved in the previous section.
Note that (\ref{F4X}) also respects hidden SO(6)$\simeq$ SU(4) invariance, with the SU(4)$/$U(3) transformations acting as
\be
\delta \varphi^i = \varepsilon^{ikl}\lambda_k\bar\varphi_l\,, \quad
\delta \bar\varphi_i = \varepsilon_{ikl}\bar\lambda^k\varphi^l\,,
\ee
where $\lambda_i$ comprise 6 corresponding group parameters. This is an indication that the superfield effective action (\ref{G}), besides
the superconformal SU(2,2$|$3) symmetry, enjoys as well an on-shell SU(4) symmetry, and hence, the
superconformal SU(2,2$|$4) symmetry as a closure of the two former ones. It would be interesting to explicitly find the realization
of this SU(4) on the analytic superfield strengths.
\subsection{$F^6/X^8$ term}
It is known that the non-conformal action (\ref{S4}) produces
not only the $F^4$ term but also the $F^6$ term in
the Born-Infeld action \cite{IZ}. The $F^6$ term appears
essentially on shell, when some of the auxiliary fields are
eliminated by their effective equations of motion. In \cite{BISZ}
it was conjectured that this procedure should work in a similar way in the
superconformal case, with the $F^6/X^8$ term as the outcome. Here
we show that this is indeed the case and present details of the relevant derivation.
As shown in \cite{IZ}, the Maxwell field strength $F_{mn}=\partial_m A_n-\partial_n A_m$
is accompanied by the antisymmetric tensor auxiliary field $H_{mn}=-H_{nm}$.
When the on-shell constraints are relaxed, these fields appear in the superfield
strengths in the combination $V_{mn}=\frac14(F_{mn}+H_{mn})$,
\be
\bar\omega^{12}=u^1_i \phi^i+4i\theta_2^\alpha\theta_3^\beta
V_{\alpha\beta}+\ldots\,,\qquad
\omega_{23}=\bar u_3^i
\bar\phi_i+4i\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}\bar
V_{\dot\alpha\dot\beta}+\ldots\,.
\label{W+aux}
\ee
Here $V_{\alpha\beta}$ and $\bar V_{\dot\alpha\dot\beta}$ are
spinorial components of the antisymmetric tensor $V_{mn}$ and dots
stand for the other field components which are irrelevant for
our consideration.
The part of the free classical action $S_2$ which involves
these fields reads \cite{IZ}
\be
S_2=\int d^4x[V^2+\bar V^2-2(VF+\bar V \bar F)+\frac12(F^2+\bar
F^2)]\,.
\label{S2}
\ee
One can recover the standard Maxwell action for
$F_{mn}$ upon eliminating the auxiliary fields from $S_2$.
However, our purpose is to eliminate them from the {\it effective}
equations of motion, when the action (\ref{G}) is added to the
classical free SYM action. The superfield effective equations
of motion are derived in Appendix D. Here we need only some SU(3) singlet
sub-sector of the component expansion of these equations, so
it is simpler to derive it independently.
As in the previous subsection, we substitute the
superfields (\ref{W+aux}) into (\ref{G}) and find
\be
\Gamma_{F^4/X^4}=\alpha\int d^4x\frac{V^2\bar
V^2}{(\varphi^i\bar\varphi_i)^2}\,.
\ee
The action $S_2+\Gamma_{F^4/X^4}$ produces the following equations
of motion for the auxiliary fields $V_{\alpha\beta}$ and
$\bar V_{\dot\alpha\dot\beta}$,
\be
F_{\alpha\beta}=V_{\alpha\beta}\left[1+\alpha\frac{\bar V^2}{(\varphi^i\bar \varphi_i)^2}
\right],\qquad
\bar F_{\dot\alpha\dot\beta}=\bar V_{\dot\alpha\dot\beta}
\left[1+\alpha\frac{ V^2}{(\varphi^i\bar \varphi_i)^2}
\right].
\ee
The solution of these equations can be represented as a series
in the Maxwell field strength, in which we need only the lowest
terms,
\be
V_{\alpha\beta}=F_{\alpha\beta}\left[
1-\alpha\frac{\bar F^2}{(\varphi^i\bar \varphi_i)^2}+O(F^3)
\right],\qquad
\bar V_{\dot\alpha\dot\beta}=\bar F_{\dot\alpha\dot\beta}\left[
1-\alpha\frac{F^2}{(\varphi^i\bar \varphi_i)^2}+O(F^3)
\right].
\ee
Substituting these solutions back into $S_2+\Gamma_{F^4/X^4}$, we earn
the correct $F^6$ term,
\be
S_2+\Gamma_{F^4/X^4}=\int d^4x\left[
-\frac12(F^2+\bar F^2)+\alpha\frac{F^2\bar F^2}{(\varphi^i\bar\varphi_i)^2}
-\alpha^2\frac{F^2\bar F^2}{(\varphi^i\bar\varphi_i)^2}(F^2+\bar F^2)
+O(F^8)
\right].
\ee
With $\alpha=Q/2$ this action coincides, up to the $F^6$ order, with
the Born-Infeld part of the effective worldvolume action for a D3 brane moving in curved
$AdS_5\times S^5$ vacuum background of type IIB supergravity
\cite{MT}\footnote{We omit here all terms with derivatives of scalars $X^I$.},
\bea
&&\int d^4x\,\frac{|X|^4}Q\left[1-
\sqrt{-\det(\eta_{mn}+Q^{1/2}|X|^{-2}F_{mn})}
\right]
\nn\\&=&
-\frac12\int d^4x[F^2+\bar F^2-\frac{Q}{|X|^4}F^2\bar F^2
+\frac12\frac{Q^2}{|X|^8}F^2\bar F^2(F^2+\bar F^2)+O(F^8)]
\,.
\eea
In stringy language, we can make the identification $Q= N g_s\alpha'{}^2/\pi$, where $N$ is
the number of D3 branes which induce the $AdS_5\times S^5$
geometry, $g_s$ is the string coupling and $\alpha'$ is the
inverse string tension. It was conjectured in
\cite{CT,M,KK,Alwis,BGL} (see also \cite{BI-review} for a review)
that such D3 brane effective action should coincide with
the low-energy effective action of $\cN=4$ SU($N$) SYM theory in the
large $N$ limit. Thus here we proved this conjecture up to the $F^6/X^8$
order.
\subsection{A comment on the Wess-Zumino term}
\label{WZ-comment}
As shown in \cite{TZ,Intriligator}, the quantum effective action of ${\cal N}=4$ SYM theory must contain a Wess-Zumino (WZ)
type non-tensor term in the scalar fields sector. The presence of such term in various superfield versions of the ${\cal N}=4$ SYM
effective action
(in particular, in its ${\cal N}=2$ superfield version \cite{BuIv}) was recently proved in \cite{BS1,BS2}. Here we give an evidence
that the ${\cal N}=3$ effective action (\ref{G}) also contains WZ term in its component expansion.
To detect the WZ term, it is sufficient to keep only scalar fields in the superfields
(\ref{W-comp}),
\bea
\hat\omega_{23}&=&\phi^1
+2i\theta_2^\alpha\bar\theta^{2\dot\alpha}\partial_{\alpha\dot\alpha}\phi^1
-4i\theta_2^\alpha\bar\theta^{1\dot\alpha}\partial_{\alpha\dot\alpha}\phi^2
-4i\theta_3^\alpha\bar\theta^{1\dot\alpha}\partial_{\alpha\dot\alpha}\phi^3
+8\theta_2^\alpha\theta_3^\beta\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}
\partial_{\alpha\dot\alpha}\partial_{\beta\dot\beta}\phi^3\,,
\nn\\
\hat{\bar \omega}{}^{12}&=&\bar\phi_3
-2i\theta_2^\alpha\bar\theta^{2\dot\alpha}\partial_{\alpha\dot\alpha}\bar\phi_3
+4i\theta_3^\alpha\bar\theta^{1\dot\alpha}\partial_{\alpha\dot\alpha}\bar\phi_1
+4i\theta_3^\alpha\bar\theta^{2\dot\alpha}\partial_{\alpha\dot\alpha}\bar\phi_2
+8\bar\theta^{1\dot\alpha}\bar\theta^{2\dot\beta}
\theta_2^\alpha\theta_3^\beta
\partial_{\alpha\dot\alpha}\partial_{\beta\dot\beta}\bar\varphi_1
\,.\nn\\&&
\eea
We substitute these superfields into the action (\ref{G}) and
integrate over the Grassmann variables, keeping only those terms which contain
four derivatives contracted with the antisymmetric $\varepsilon$-symbol,
\bea
\Gamma_{WZ}&=&-\frac{i\alpha}{8}\varepsilon^{mnpq}\int d^4x du
[\partial_m\phi^2 \partial_n\bar\phi_3\partial_p\bar\phi_2\partial_q\phi^3
+\partial_m\bar\phi_2\partial_n\bar\phi_1\partial_p\phi^2\partial_q\phi^1]
\nn\\&&\times
\sum_{i,j=0}^\infty (-1)^{i+j}\frac{(i+j+2)!(i+1)(j+1)}{i!j!}
(c^3\bar\phi_3)^i(\bar c_1\phi^1)^j
\,.
\label{WZ-series}
\eea
To compare this expression with the standard expression (\ref{WZ}) for WZ
term\footnote{To be precise, we compare
(\ref{WZ-series}) with the WZ action in the
four-dimensional form (\ref{SWZ}).}, it is necessary to compute the harmonic integrals and to
sum the series. Unfortunately, it is very difficult to find the
explicit expression for the integral
\be
\int du\,
u^1_{i_1} \bar u_1^{i'_1}\ldots
u^1_{i_n} \bar u_1^{i'_n}
u^3_{j_1} \bar u_3^{j'_1}\ldots
u^3_{j_m} \bar u_3^{j'_m}u^2_k \bar u_2^{k'}
\ee
in terms of (anti)symmetrized irreducible combinations of
the delta-symbols. Therefore here we restrict ourselves to considering only
the lowest terms in (\ref{WZ-series}), namely,
\be
\Gamma_{WZ}=\frac32i\alpha\varepsilon^{mnpq}\int d^4x du
[\partial_m\phi^2 \partial_n\bar\phi_3\partial_p\bar\phi_2\partial_q\phi^3
+\partial_m\bar\phi_2\partial_n\bar\phi_1\partial_p\phi^2\partial_q\phi^1]
(c^3\bar\phi_3+\bar c_1\phi^1)
+O(\phi^6)
\,.
\label{WZ-leading}
\ee
The corresponding harmonic integral is quite easy to do,
\be
\int du\,u^1_iu^2_ju^3_k \bar u_1^{i'} \bar u_2^{j'} \bar u_3^{k'}
=\frac1{36}\varepsilon_{ijk}\varepsilon^{i'j'k'}
+\frac1{60}\delta_i^{(i'}\delta_j^{j'}\delta_k^{k')}
+\frac1{18}\delta_i^{(i'}\delta_j^{[j')}\delta_k^{k']}
+\frac1{18}\delta_i^{[i'}\delta_j^{(j']}\delta_k^{k')}
\,.
\label{25}
\ee
Then it is straightforward to see that only the first term in the r.h.s. of (\ref{25})
contributes to (\ref{WZ-leading}), while all other terms either vanish
after contracting the indices or form total derivatives. As a result, (\ref{WZ-leading})
can be rewritten as
\bea
\Gamma_{WZ}&=&\frac{i\alpha}{24}\varepsilon^{mnpq}\int d^4x\,
\varepsilon_{ijk}\varepsilon^{i'j'k'}
[c^i\partial_m\phi^j \partial_n \phi^k
\bar\phi_{i'} \partial_p\bar\phi_{j'}\partial_q\bar\phi_{k'}
\nn\\&&\qquad\qquad
-\,\phi^i \partial_m\phi^j\partial_n\phi^k
\bar c_{i'}\partial_p\bar \phi_{j'}\partial_q\bar\phi_{k'}]
+O(\phi^6)\,.
\eea
This expression coincides with (\ref{25_}) under the choice
\be
\alpha=-\frac1{2\pi^2}\,.
\ee
This proves that the action (\ref{G}) contains the Wess-Zumino
term. One of the possible four-dimensional representations of this term is given
by the expression (\ref{SWZ}).
\section{Summary and discussion}
In the present paper we made an essential step towards solving the
long-standing problem of constructing $\cN=3$ SYM low-energy effective action in
terms of unconstrained $\cN=3$ superfields. We constructed the
leading part of this effective action which is responsible for the
$F^4/X^4$ term in components. This action is given by a
local functional in the $\cN=3$ analytic superspace, such that it depends on
the $\cN=3$ superfield strength without derivatives on them. The form of this functional is
uniquely fixed by the requirements of scale and $\gamma_5$ invariance,
although the action respects further SU(2,2$|$3) superconformal symmetry
(and, perhaps, SU(2,2$|$4)).
Since the $\cN=3$ and $\cN=4$ SYM models are equivalent on shell,
the action (\ref{G}) provides us with an $\cN=3$ superfield
description of the $\cN=4$ SYM low-energy effective action.
This effective action was previously studied in the $\cN=2$
harmonic superspace \cite{BuIv,BIP} and was rewritten in terms of the
on-shell $\cN=4$
superfields in \cite{BLS,BS1,BS2}. In contrast to the representations
of this action in the $\cN=2$ and $\cN=4$ harmonic superspaces,
the Lagrangian in (\ref{G}) has an explicit dependence on the vev
constants $c^i=\langle\varphi^i\rangle$. However, this dependence
is rather spurious: we proved that the action itself is in fact independent of any
particular choice of these constants. This phenomenon is very similar to what one observed
in the action of the $\cN=2$ improved tensor multiplet \cite{deWit} in the harmonic
superspace \cite{GIO} which also explicitly included the vev constants of the scalars,
but this dependence disappeared in the full component action. In \cite{GIO} it was argued that
the presence of such constants reflects the non-trivial topological properties of
this action. In our case the $\cN=4$ SYM low-energy
effective action also contains some topological term given by the
WZ action for the scalar fields \cite{TZ,Intriligator}.
Therefore the action (\ref{G}) can be equally considered as an $\cN=3$
superfield extension of the WZ term.
A possible form of the WZ term arising from the $\cN=3$ harmonic
superspace is discussed in Appendix C, see eq. (\ref{SWZ}).
The constants $c^i$ in this action break the
manifest SU(3) symmetry, though the action is still SU(3) and SU(4)
invariant up to total derivatives. This confirms the conclusions
of \cite{BS1} that the four-dimensional WZ term cannot be made manifestly
invariant under SU(3) since this group is anomalous.
In the present paper we studied the bosonic component structure of the
action (\ref{G}) in the limit of constant Maxwell and scalar
fields and argued that it contains the Wess-Zumino term.
We showed that this action correctly reproduces the coefficients in
front of the $F^4/X^4$ and $F^6/X^8$ terms to ensure their coincidence with the similar
terms in the worldvolume action of D3 brane in the $AdS_5\times S^5$ background. To make the comparison of
the action (\ref{G}) with the D3 brane action more precise, it is necessary to
study the component structure of (\ref{G}) in the scalar field
sector in more detail, beyond the constant field
approximation. This problem is technically involved
and will be addressed elsewhere.
Finally, it is worth pointing out that the action (\ref{G}) was derived
solely by employing the group-theory requirements of gauge invariance and superconformal
symmetry. It is very desirable to develop the background field
method for the $\cN=3$ SYM theory in order to re-derive the action (\ref{G})
from the quantum perturbation theory in $\cN=3$ harmonic superspace\footnote{Like as the ${\cal N}=2$ superfield effective
action of ref.\ \cite{BuIv} was re-derived from the ${\cal N}=2$ harmonic superfield perturbation theory in
\cite{BIP}.}.
Note that the free propagators in the $\cN=3$ harmonic superspace were
studied in \cite{DM}. These methods might help to unveil the
structure of effective action in the $\cN=3$ and $\cN=4$ SYM
models beyond the low-energy approximation.
\vspace{30pt}
\noindent
{\bf Acknowledgments}\\[3mm]
I.B.S.\ is indebted to D.~Belyaev, S.~Kuzenko, W.~Schulgin and D.~Sorokin for
useful discussions. I.L.B.\ is grateful to CAPES for supporting his visit to the Physics
Department of Universidade Federal de Juiz de Fora where the final
part of work was done.
The authors are grateful to the RFBR grant Nr.\ 11-02-90445 for partial support.
I.L.B.\ and I.B.S.\ acknowledge the support from the RFBR grant Nr.\ 12-02-00121 and
from LRSS grant Nr.\ 224.2012.2.
The work of I.B.S.\ was also supported by the Marie Curie research fellowship Nr.\ 236231,
``QuantumSupersymmetry''. E.A.I. and B.M.Z. acknowledge the support from
the RFBR grant Nr.\ 09-02-01209 and a grant of Heisenberg-Landau Program.
E.A.I.\ thanks the Directorate of SUBATECH, University of Nantes, for the kind hospitality at the final stage
of this work.
|
\section{Introduction, Definitions and Preliminaries}
Throughout this paper, we use the following standard notations:
$\mathbb{N}=\{1,2,3,$\ldots $\}$, $\mathbb{N}_{0}=\{0,1,2,3,$\ldots $\}
\mathbb{N}\cup \{0\}$ and $\mathbb{Z}^{-}=\{-1,-2,-3,$\ldots $\}$. Here,
\mathbb{Z}$ denotes the set of integers, $\mathbb{R}$ denotes the set of
real numbers and $\mathbb{C}$ denotes the set of complex numbers. We assume
that $\ln (z)$ denotes the principal branch of the multi-valued function
\ln (z)$ with the imaginary part $\Im \left( \ln (z)\right) $ constrained b
\begin{equation*}
-\pi <\Im \left( \ln (z)\right) \leq \pi .
\end{equation*
Furthermore,
\begin{equation*}
0^{n}=\left\{
\begin{array}{cc}
1 & n=0 \\
& \\
0 & n\in \mathbb{N}
\end{array
\right.
\end{equation*
\begin{equation*}
\left(
\begin{array}{c}
x \\
\end{array
\right) =\frac{x(x-1)\cdots (x-v+1)}{v!}
\end{equation*
an
\begin{equation*}
\left\{ z\right\} _{0}=1\text{ and }\left\{ z\right\}
_{j}=\dprod\limits_{d=0}^{j-1}(z-d),
\end{equation*
where $j\in \mathbb{N}$ and $z\in \mathbb{C}$ cf. (\cite{Comtet}, \cit
{LuoSrivatava2010}).
The generating functions have various applications in many branches of
Mathematics and Mathematical Physics. These functions are defined by linear
polynomials, differential relations, globally referred to as \textit
functional equations}. The functional equations arise in well-defined
combinatorial contexts and they lead systematically to well-defined classes
of functions (cf. see, for detail, \cite{Flajolet}). Although, in the
literature, one can find extensive investigations related to the generating
functions for the Bernoulli, Euler and Genocchi numbers and polynomials and
also their generalizations, the $\lambda $-Stirling numbers of the second
kind, the array polynomials and the Eulerian polynomials, related to
nonnegative real parameters, have not been studied yet. Therefore, Section
2, Section 3 and Section 4 of this paper deal with new classes of generating
functions which are related to generalized $\lambda $-Stirling type numbers
of the second kind, generalized array type polynomials and generalized
Eulerian polynomials, respectively. By using these generating functions, we
derive many functional equations and differential equations. By using these
equations, we investigate and introduce fundamental properties and many new
identities for the generalized $\lambda $-Stirling type numbers of the
second kind, the generalized array type polynomials and the generalized
Eulerian type polynomials and numbers. We also derive multiplication
formulas and recurrence relations for these numbers and polynomials.
The remainder of this study is organized as follows:
In section 5, we derive new identities related to the generalized Bernoulli
polynomials, the generalized Eulerian type polynomials, generalized $\lambda
$-Stirling type numbers and the generalized array polynomials.
In section 6, we give relations between generalized Bernoulli polynomials
and generalized array polynomials.
In section 7, We give an application of the Laplace transform to the
generating functions for the generalized Bernoulli polynomials and the
generalized array type polynomials.
In section 8, by using the bosonic and the fermionic $p$-adic integral on
\mathbb{Z}_{p}$, we find some new identities related to the Bernoulli
polynomials, the generalized Eulerian type polynomials and Stirling numbers.
\section{Generating Function for generalized $\protect\lambda $-Stirling
type numbers of the second kind}
The Stirling numbers are used in combinatorics, in number theory, in
discrete probability distributions for finding higher order moments, etc.
The Stirling number of the second kind, denoted by $S(n,k)$, is the number
of ways to partition a set of $n$ objects into $k$ groups. These numbers
occur in combinatorics and in the theory of partitions.
In this section, we construct a new generating function, related to
nonnegative real parameters, for the generalized $\lambda $-Stirling type
numbers of the second kind. We derive some elementary properties including
recurrence relations of these numbers. The following definition provides a
natural generalization and unification of the $\lambda $-Stirling numbers of
the second kind:
\begin{definition}
Let $a$,$~b\in \mathbb{R}^{+}$ ($a\neq b$), $\lambda \in \mathbb{C}$ and
v\in \mathbb{N}_{0}$. The generalized $\lambda $-Stirling type numbers of
the second kind $\mathcal{S}(n,v;a,b;\lambda )$\ are defined by means of the
following generating function
\begin{equation}
f_{S,v}(t;a,b;\lambda )=\frac{\left( \lambda b^{t}-a^{t}\right) ^{v}}{v!
=\sum_{n=0}^{\infty }\mathcal{S}(n,v;a,b;\lambda )\frac{t^{n}}{n!}.
\label{s1}
\end{equation}
\end{definition}
\begin{remark}
By setting $a=1$ and $b=e$ in (\ref{s1}), we have the $\lambda $-Stirling
numbers of the second kin
\begin{equation*}
\mathcal{S}(n,v;1,e;\lambda )=S(n,v;\lambda )
\end{equation*
which are defined by means of the following generating function
\begin{equation*}
\frac{\left( \lambda e^{t}-1\right) ^{v}}{v!}=\sum_{n=0}^{\infty
}S(n,v;\lambda )\frac{t^{n}}{n!},
\end{equation*
cf. (\cite{LuoSrivatava2010}, \cite{Srivastava2011}). Substituting $\lambda
=1$ into above equation, we have the Stirling numbers of the second kin
\begin{equation*}
S(n,v;1)=S(n,v),
\end{equation*
cf. (\cite{Comtet}, \cite{LuoSrivatava2010}, \cite{Srivastava2011}). These
numbers have the following well known properties
\begin{equation*}
S(n,0)=\delta _{n,0},
\end{equation*
\begin{equation*}
S(n,1)=S(n,n)=1
\end{equation*
an
\begin{equation*}
S(n,n-1)=\left(
\begin{array}{c}
n \\
\end{array
\right) ,
\end{equation*
where $\delta _{n,0}$ denotes the Kronecker symbol (see \cite{Comtet}, \cit
{LuoSrivatava2010}, \cite{Srivastava2011}).
\end{remark}
By using (\ref{s1}), we obtain the following theorem:
\begin{theorem}
\label{Theorem STnumber
\begin{equation}
\mathcal{S}(n,v;a,b;\lambda )=\frac{1}{v!}\sum_{j=0}^{v}(-1)^{j}\left(
\begin{array}{c}
v \\
\end{array
\right) \lambda ^{v-j}\left( j\ln a+(v-j)\ln b\right) ^{n} \label{as1}
\end{equation
an
\begin{equation}
\mathcal{S}(n,v;a,b;\lambda )=\frac{1}{v!}\sum_{j=0}^{v}(-1)^{v-j}\left(
\begin{array}{c}
v \\
\end{array
\right) \lambda ^{j}\left( j\ln b+(v-j)\ln a\right) ^{n}. \label{as1a}
\end{equation}
\end{theorem}
\begin{proof}
By using (\ref{s1}) and the binomial theorem, we can easily arrive at the
desired results.
\end{proof}
By using the formula (\ref{as1}), we can compute some values of the numbers
\mathcal{S}(n,v;a,b;\lambda )$ as follows
\begin{equation*}
\mathcal{S}(0,0;a,b;\lambda )=1,
\end{equation*
\begin{equation*}
\mathcal{S}(0,0;a,b;\lambda )=1,
\end{equation*
\begin{equation*}
\mathcal{S}(1,0;a,b;\lambda )=0,
\end{equation*
\begin{equation*}
\mathcal{S}(1,1;a,b;\lambda )=\ln \left( \frac{b^{\lambda }}{a}\right) ,
\end{equation*
\begin{equation*}
\mathcal{S}(2,0;a,b;\lambda )=0,
\end{equation*
\begin{equation*}
\mathcal{S}(2,1;a,b;\lambda )=\lambda \left( \ln b\right) ^{2}-\left( \ln
a\right) ^{2},
\end{equation*
\begin{equation*}
\mathcal{S}(2,2;a,b;\lambda )=\frac{\lambda ^{2}}{2}\left( \ln b^{2}\right)
^{2}-\lambda \ln \left( ab\right) +\left( \ln a^{2}\right) ^{2},
\end{equation*}
\begin{equation*}
\mathcal{S}(3,0;a,b;\lambda )=0,
\end{equation*
\begin{equation*}
\mathcal{S}(3,1;a,b;\lambda )=\lambda \left( \ln b\right) ^{3}-\left( \ln
a\right) ^{3},
\end{equation*
\begin{equation*}
\mathcal{S}(0,v;a,b;\lambda )=\frac{\left( \lambda -1\right) ^{v}}{v!},
\end{equation*
\begin{equation*}
\mathcal{S}(n,0;a,b;\lambda )=\delta _{n,0}
\end{equation*
an
\begin{equation*}
\mathcal{S}(n,1;a,b;\lambda )=\lambda \left( \ln b\right) ^{n}-\left( \ln
a\right) ^{n}.
\end{equation*}
\begin{remark}
By setting $a=1$ and $b=e$ in the assertions (\ref{as1}) of Theorem \re
{Theorem STnumber}, we have the following result
\begin{equation*}
S(n,v;\lambda )=\frac{1}{v!}\sum_{j=0}^{v}\left(
\begin{array}{c}
v \\
\end{array
\right) \lambda ^{v-j}(-1)^{j}\left( v-j\right) ^{n}.
\end{equation*
The above relation has been studied by Srivastava \cite{Srivastava2011} and
Luo \cite{LuoSrivatava2010}. By setting $\lambda =1$ in the above equation,
we have the following result
\begin{equation*}
S(n,v;\lambda )=\frac{1}{v!}\sum_{j=0}^{v}\left(
\begin{array}{c}
v \\
\end{array
\right) (-1)^{j}\left( v-j\right) ^{n}
\end{equation*
cf. (\cite{AgohDilcher}, \cite{cagic}, \cite{Carlitz}, \cite{Carlitz1953G},
\cite{Comtet}, \cite{T. Kim}, \cite{LuoSrivatava2010}, \cite{SimsekSpringer
, \cite{YsimsekStirling}, \cite{Srivastava2011}, \cite{SrivastawaGargeSC}).
\end{remark}
By differentiating both sides of equation (\ref{s1}) with respect to the
variable $t$, we obtain the following\textit{\ }differential equations
\begin{equation*}
\frac{\partial }{\partial t}f_{S,v}(t;a,b;\lambda )=\left( \lambda (\ln
b)b^{t}-(\ln a)a^{t}\right) f_{S,v-1}(t;a,b;\lambda )
\end{equation*
o
\begin{equation}
\frac{\partial }{\partial t}f_{S,v}(t;a,b;\lambda )=v\ln
(b)f_{S,v}(t;a,b;\lambda )+\ln \left( \frac{b}{a}\right)
a^{t}f_{S,v-1}(t;a,b;\lambda ). \label{s1a}
\end{equation}
By using equations (\ref{s1}) and (\ref{s1a}), we obtain recurrence
relations for the generalized $\lambda $-Stirling type numbers of the second
kind by the following theorem:
\begin{theorem}
\label{TE2} Let $n,v\in \mathbb{N}$
\begin{equation}
\mathcal{S}(n,v;a,b;\lambda )=\sum_{j=0}^{n-1}\left(
\begin{array}{c}
n-1 \\
\end{array
\right) \mathcal{S}(j,v-1;a,b;\lambda )\left( \lambda \left( \ln (b)\right)
^{n-j}-\left( \ln (a)\right) ^{n-j}\right) . \label{s4}
\end{equation
o
\begin{eqnarray*}
\mathcal{S}(n,v;a,b;\lambda ) &=&v\ln (b)\mathcal{S}(n-1,v;a,b;\lambda ) \\
&&+\ln \left( \frac{b}{a}\right) \sum_{j=0}^{n-1}\left(
\begin{array}{c}
n-1 \\
\end{array
\right) \mathcal{S}(j,v-1;a,b;\lambda )\left( \ln (a)\right) ^{n-1-j}.
\end{eqnarray*}
\end{theorem}
\begin{remark}
By setting $a=1$ and $b=e$, Theorem \ref{TE2} yields the corresponding
results which are proven by Luo and Srivastava \cite[Theorem 11
{LuoSrivatava2010}. Substituting $a=\lambda =1$ and $b=e$ into Theorem \re
{TE2}, we obtain the following known results
\begin{equation*}
S(n,v)=\sum_{j=0}^{n-1}\left(
\begin{array}{c}
n-1 \\
\end{array
\right) S(j,v-1),
\end{equation*
an
\begin{equation*}
S(n,v)=vS(n-1,v)+S(n-1,v-1),
\end{equation*
cf. (\cite{AgohDilcher}, \cite{Carlitz1976}, \cite{Comtet}, \cit
{LuoSrivatava2010}, \cite{SimsekSpringer}, \cite{YsimsekStirling}).
\end{remark}
The generalized $\lambda $-Stirling type numbers of the second kind can also
be defined by equation (\ref{s5}):
\begin{theorem}
\label{T3}Let $k\in \mathbb{N}_{0}$ and $\lambda \in \mathbb{C}$
\begin{equation}
\lambda ^{x}\left( \ln b^{x}\right) ^{m}=\sum_{l=0}^{m}\sum_{j=0}^{\infty
}\left(
\begin{array}{c}
m \\
\end{array
\right) \left(
\begin{array}{c}
x \\
\end{array
\right) j!\mathcal{S}(l,j;a,b;\lambda )\left( \ln \left( a^{(x-j)}\right)
\right) ^{m-l}. \label{s5}
\end{equation}
\end{theorem}
\begin{proof}
By using (\ref{s1}), we ge
\begin{equation*}
\left( \lambda b^{t}\right) ^{x}=\sum_{j=0}^{\infty }\left(
\begin{array}{c}
x \\
\end{array
\right) j!\sum_{m=0}^{\infty }\mathcal{S}(m,j;a,b;\lambda )\frac{t^{m}}{m!
\sum_{n=0}^{\infty }(\ln a^{x-j})^{n}\frac{t^{n}}{n!}.
\end{equation*
From the above equation, we obtai
\begin{equation*}
\lambda ^{x}\sum_{m=0}^{\infty }\left( \ln b\right) ^{m}\frac{t^{m}}{m!
=\sum_{m=0}^{\infty }\sum_{j=0}^{\infty }\left(
\begin{array}{c}
x \\
\end{array
\right) j!\mathcal{S}(m,j;a,b;\lambda )\frac{t^{m}}{m!}\sum_{n=0}^{\infty
}(\ln a^{x-j})^{n}\frac{t^{n}}{n!}.
\end{equation*
Therefor
\begin{equation*}
\lambda ^{x}\sum_{m=0}^{\infty }\left( \ln b\right) ^{m}\frac{t^{m}}{m!
=\sum_{m=0}^{\infty }\left( \sum_{l=0}^{m}\sum_{j=0}^{\infty }\left(
\begin{array}{c}
m \\
\end{array
\right) \left(
\begin{array}{c}
x \\
\end{array
\right) j!\mathcal{S}(l,j;a,b;\lambda )\left( \ln a^{(x-j)}\right)
^{m-l}\right) \frac{t^{m}}{m!}.
\end{equation*
Comparing the coefficients of $\frac{t^{m}}{m!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
For $a=0$ and $b=e$, the formula (\ref{s5}) can easily be shown to be
reduced to the following result which is given by Luo and Srivastava \cite
Theorem 9]{LuoSrivatava2010}
\begin{equation*}
\lambda ^{x}x^{n}=\sum_{l=0}^{\infty }\left(
\begin{array}{c}
x \\
\end{array
\right) l!S(n,l;\lambda ),
\end{equation*
where $n\in \mathbb{N}_{0}$ and $\lambda \in \mathbb{C}$. For $\lambda =1$,
the above formula is reduced t
\begin{equation*}
x^{n}=\sum_{v=0}^{n}\left(
\begin{array}{c}
x \\
\end{array
\right) v!S(n,v)
\end{equation*
cf. (\cite{AgohDilcher}, \cite{Carlitz1976}, \cite{Comtet}, \cite{T. Kim},
\cite{LuoSrivatava2010}).
\end{remark}
\section{Generalized array type polynomials}
By using the same motivation with the $\lambda $-Stirling type numbers of
the second kind, we also construct a novel generating function, related to
nonnegative real parameters, of the \textit{generalized array type
polynomials}. We derive some elementary properties including recurrence
relations of these polynomials. The following definition provides a natural
generalization and unification of the array polynomials:
\begin{definition}
Let $a$, $b\in \mathbb{R}^{+}$ ($a\neq b$), $x\in \mathbb{R}$, $\lambda \in
\mathbb{C}$ and $v\in \mathbb{N}_{0}$. The generalized array type
polynomials $\mathcal{S}_{v}^{n}(x;a,b;\lambda )$\ can be defined b
\begin{equation}
\mathcal{S}_{v}^{n}(x;a,b;\lambda )=\frac{1}{v!}\sum_{j=0}^{v}(-1)^{v-j
\left(
\begin{array}{c}
v \\
\end{array
\right) \lambda ^{j}\left( \ln \left( a^{v-j}b^{x+j}\right) \right) ^{n}.
\label{as2}
\end{equation}
\end{definition}
By using the formula (\ref{as2}), we can compute some values of the
polynomials $\mathcal{S}_{v}^{n}(x;a,b;\lambda )$ as follows
\begin{equation*}
\mathcal{S}_{0}^{n}(x;a,b;\lambda )=\left( \ln \left( b^{x}\right) \right)
^{n},
\end{equation*
\begin{equation*}
\mathcal{S}_{v}^{0}(x;a,b;\lambda )=\frac{\left( 1-\lambda \right) ^{v}}{v!}
\end{equation*
an
\begin{equation*}
\mathcal{S}_{1}^{1}(x;a,b;\lambda )=-\ln (ab^{x})+\lambda \ln (b^{x+1}).
\end{equation*}
\begin{remark}
The polynomials $\mathcal{S}_{v}^{n}(x;a,b;\lambda )$ may be also called
generalized $\lambda $-array type polynomials. By substituting $x=0$ into
\ref{as2}), we arrive at (\ref{as1a})
\begin{equation*}
\mathcal{S}_{v}^{n}(0;a,b;\lambda )=\mathcal{S}(n,v;a,b;\lambda ).
\end{equation*
Setting $a=\lambda =1$ and $b=e$ in (\ref{as2}), we hav
\begin{equation*}
S_{v}^{n}(x)=\frac{1}{v!}\sum_{j=0}^{v}(-1)^{v-j}\left(
\begin{array}{c}
v \\
\end{array
\right) \left( x+j\right) ^{n},
\end{equation*
a result due to Chang and Ha \cite[Eq-(3.1)]{Chan}, Simsek \cit
{SimsekSpringer}. It is easy to see tha
\begin{equation*}
S_{0}^{0}(x)=S_{n}^{n}(x)=1,
\end{equation*
\begin{equation*}
S_{0}^{n}(x)=x^{n}
\end{equation*
and for $v>n$
\begin{equation*}
S_{v}^{n}(x)=0
\end{equation*
cf. \cite[Eq-(3.1)]{Chan}.
\end{remark}
Generating functions for the polynomial $\mathcal{S}_{v}^{n}(x;a,b,c;\lambda
)$ can be defined as follows:
\begin{definition}
Let $a$, $b\in \mathbb{R}^{+}$ ($a\neq b$), $\lambda \in \mathbb{C}$ and
v\in \mathbb{N}_{0}$. The generalized array type polynomials $\mathcal{S
_{v}^{n}(x;a,b;\lambda )$\ are defined by means of the following generating
function
\begin{equation}
g_{v}(x,t;a,b;\lambda )=\sum_{n=0}^{\infty }\mathcal{S}_{v}^{n}(x;a,b
\lambda )\frac{t^{n}}{n!}. \label{ab1}
\end{equation}
\end{definition}
\begin{theorem}
Let $a$, $b\in \mathbb{R}^{+}$, ($a\neq b$), $\lambda \in \mathbb{C}$ and
v\in \mathbb{N}_{0}$
\begin{equation}
g_{v}(x,t;a,b;\lambda )=\frac{1}{v!}\left( \lambda b^{t}-a^{t}\right)
^{v}b^{xt}. \label{ab0}
\end{equation}
\end{theorem}
\begin{proof}
By substituting (\ref{as2}) into the right hand side of (\ref{ab1}), we
obtai
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{S}_{v}^{n}(x;a,b;\lambda )\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\left( \frac{1}{v!}\sum_{j=0}^{v}(-1)^{v-j}\left(
\begin{array}{c}
v \\
\end{array
\right) \lambda ^{j}\left( \ln \left( a^{v-j}b^{x+j}\right) \right)
^{n}\right) \frac{t^{n}}{n!}.
\end{equation*
Therefor
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{S}_{v}^{n}(x;a,b;\lambda )\frac{t^{n}}{n!}
\frac{1}{v!}\sum_{j=0}^{v}(-1)^{v-j}\left(
\begin{array}{c}
v \\
\end{array
\right) \lambda ^{j}\sum_{n=0}^{\infty }\left( \ln \left(
a^{v-j}b^{x+j}\right) \right) ^{n}\frac{t^{n}}{n!}.
\end{equation*
The right hand side of the above equation is the Taylor series for $e^{(\ln
\left( a^{v-j}b^{x+j}\right) )t}$, thus we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{S}_{v}^{n}(x;a,b;\lambda )\frac{t^{n}}{n!
=\left( \frac{1}{v!}\sum_{j=0}^{v}(-1)^{v-j}\left(
\begin{array}{c}
v \\
\end{array
\right) \lambda ^{j}a^{\left( v-j\right) t}b^{jt}\right) b^{xt}.
\end{equation*}
By using (\ref{s1}) and binomial theorem in the above equation, we arrive at
the desired result.
\end{proof}
\begin{remark}
If we set $\lambda =1$ in (\ref{ab0}), we arrive a new special case of the
array polynomials given b
\begin{equation*}
f_{S,v}(t;a,b)b^{tx}=\sum_{n=0}^{\infty }\mathcal{S}_{v}^{n}(x;a,b)\frac
t^{n}}{n!}.
\end{equation*
In the special case whe
\begin{equation*}
a=\lambda =1\text{ and }b=e,
\end{equation*
the generalized array polynomials $\mathcal{S}_{v}^{n}(x;a,b;\lambda )$
defined by (\ref{ab0}) would lead us at once to the classical array
polynomials $\mathcal{S}_{v}^{n}(x)$, which are defined by means of the
following generating function
\begin{equation*}
\frac{\left( e^{t}-1\right) ^{v}}{v!}e^{tx}=\sum_{n=0}^{\infty }S_{v}^{n}(x
\frac{t^{n}}{n!},
\end{equation*
which yields to the generating function for the array polynomials
S_{v}^{n}(x)$ studied by Chang and Ha \cite{Chan} see also cf. (\cite{cagic
, \cite{SimsekSpringer}).
\end{remark}
The polynomials $\mathcal{S}_{v}^{n}(x;a,b;\lambda )$\ defined by (\ref{ab0
) have many interesting properties which we give in this section.
We se
\begin{equation}
g_{v}(x,t;a,b;\lambda )=b^{xt}f_{S,v}(t;a,b;\lambda ). \label{1Sse}
\end{equation}
\begin{theorem}
The following formula holds true
\begin{equation}
\mathcal{S}_{v}^{n}(x;a,b;\lambda )=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{S}(j,v;a,b;\lambda )\left( \ln b^{x}\right) ^{n-j}.
\label{1Ssc}
\end{equation}
\end{theorem}
\begin{proof}
By using (\ref{1Sse}), we obtai
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{S}_{v}^{n}(x;a,b;\lambda )\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\mathcal{S}(n,v;a,b;\lambda )\frac{t^{n}}{n!
\sum_{n=0}^{\infty }\left( \ln b^{x}\right) ^{n}\frac{t^{n}}{n!}.
\end{equation*}
From the above equation, we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{S}_{v}^{n}(x;a,b;\lambda )\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\left( \sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{S}(j,v;a,b)\left( \ln b^{x}\right) ^{n-j}\right) \frac{t^{n
}{n!}.
\end{equation*}
Comparing the coefficients of $t^{n}$ on both sides of the above equation,
we arrive at the desired result.
\end{proof}
\begin{remark}
In the special case when $a=\lambda =1$ and $b=e$, equation (\ref{1Ssc}) is
reduced t
\begin{equation*}
S_{v}^{n}(x)=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) x^{n-j}S(j,v)
\end{equation*
cf. \cite[Theorem 2]{SimsekSpringer}.
\end{remark}
By differentiating $j$ times both sides of (\ref{ab0}) with respect to the
variable $x$, we obtain the following differential equation
\begin{equation*}
\frac{\partial ^{j}}{\partial x^{j}}g_{v}(x,t;a,b;\lambda )=t^{j}\left( \ln
b\right) ^{j}g_{v}(x,t;a,b;\lambda ).
\end{equation*}
From this equation, we arrive at higher order derivative of \ the array type
polynomials by the following theorem:
\begin{theorem}
\label{TEo2} Let $n$, $j\in \mathbb{N}$ with $j\leq n$. Then we hav
\begin{equation*}
\frac{\partial ^{j}}{\partial x^{j}}\mathcal{S}_{v}^{n}(x;a,b;\lambda
)=\left\{ n\right\} _{j}\left( \ln (b)\right) ^{j}\mathcal{S
_{v}^{n-j}(x;a,b;\lambda ).
\end{equation*}
\end{theorem}
\begin{remark}
By setting $a=\lambda =j=1$ and $b=e$ in Theorem \ref{TEo2}, we hav
\begin{equation*}
\frac{d}{dx}S_{v}^{n}(x)=nS_{v}^{n-1}(x)
\end{equation*
cf. \cite{SimsekSpringer}.
\end{remark}
From (\ref{ab0}), we get the following functional equation
\begin{equation}
g_{v_{1}}(x_{1},t;a,b;\lambda )g_{v_{2}}(x_{2},t;a,b;\lambda )=\left(
\begin{array}{c}
v_{1}+v_{2} \\
v_{1
\end{array
\right) g_{v_{1}+v_{2}}(x_{1}+x_{2},t;a,b;\lambda ). \label{as3}
\end{equation
From this functional equation, we obtain the following identity:
\begin{theorem}
\begin{equation*}
\left(
\begin{array}{c}
v_{1}+v_{2} \\
v_{1
\end{array
\right) \mathcal{S}_{v_{1}+v_{2}}^{n}(x_{1}+x_{2};a,b;\lambda
)=\dsum\limits_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{S}_{v_{1}}^{j}(x_{1};a,b;\lambda )\mathcal{S
_{v_{2}}^{n-j}(x_{2};a,b;\lambda ).
\end{equation*}
\end{theorem}
\begin{proof}
Combining (\ref{ab1}) and (\ref{as3}), we ge
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\mathcal{S}_{v_{1}}^{n}(x_{1};a,b;\lambda )\frac{t^{n
}{n!}\sum_{n=0}^{\infty }\mathcal{S}_{v_{2}}^{n}(x_{2};a,b;\lambda )\frac
t^{n}}{n!} \\
&=&\left(
\begin{array}{c}
v_{1}+v_{2} \\
v_{1
\end{array
\right) \sum_{n=0}^{\infty }\mathcal{S}_{v_{1}+v_{2}}^{n}(x_{1}+x_{2};a,b
\lambda )\frac{t^{n}}{n!}.
\end{eqnarray*
Therefor
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\left( \dsum\limits_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{S}_{v_{1}}^{j}(x_{1};a,b;\lambda )\mathcal{S
_{v_{2}}^{n-j}(x_{2};a,b;\lambda )\right) \frac{t^{n}}{n!} \\
&=&\left(
\begin{array}{c}
v_{1}+v_{2} \\
v_{1
\end{array
\right) \sum_{n=0}^{\infty }\mathcal{S}_{v_{1}+v_{2}}^{n}(x_{1}+x_{2};a,b
\lambda )\frac{t^{n}}{n!}\text{.}
\end{eqnarray*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\section{Generalized Eulerian type numbers and polynomials}
In this section, we provide generating functions, related to nonnegative
real parameters, for the generalized Eulerian type polynomials and numbers,
that is, the so called \textit{generalized Apostol type Frobenius Euler
polynomials} \textit{and numbers}. We derive fundamental properties,
recurrence relations and many new identities for these polynomials and
numbers based on the generating functions, functional equations and
differential equations.
These polynomials and numbers have many applications in many branches of
Mathematics.
The following definition gives us a natural generalization of the Eulerian
polynomials:
\begin{definition}
Let $a,$ $b\in \mathbb{R}^{+}$ $(a\neq b),$ $x\in \mathbb{R},$ $\lambda \in
\mathbb{C}$ and $u\in \mathbb{C\diagdown }\left\{ 1\right\} $. The\
generalized Eulerian type polynomials $\mathcal{H}_{n}(x;u;a,b,c;\lambda )$
are defined by means of the following generating function
\begin{equation}
F_{\lambda }(t,x;u,a,b,c)=\frac{\left( a^{t}-u\right) c^{xt}}{\lambda b^{t}-
}=\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!}.
\label{4ge1}
\end{equation}
\end{definition}
By substituting $x=0$ into (\ref{4ge1}), we obtai
\begin{equation*}
\mathcal{H}_{n}(0;u;a,b,c;\lambda )=\mathcal{H}_{n}(u;a,b,c;\lambda ),
\end{equation*
where $\mathcal{H}_{n}(u;a,b,c;\lambda )$ denotes \textit{generalized
Eulerian type numbers}.
\begin{remark}
Substituting $a=1$ into (\ref{4ge1}), we hav
\begin{equation*}
\frac{\left( 1-u\right) c^{xt}}{\lambda b^{t}-u}=\sum_{n=0}^{\infty
\mathcal{H}_{n}(x;u;1,b,c;\lambda )\frac{t^{n}}{n!}
\end{equation*
a result due to Kurt and Simsek \cite{burakSimsek}. In their special case
when $\lambda =1$ and $b=c=e$, the \textit{generalized }Eulerian type
polynomials $\mathcal{H}_{n}(x;u;1,b,c;\lambda )$\ are reduced to the
Eulerian polynomials or Frobenius Euler polynomials which are defined by
means of the following generating function
\begin{equation}
\frac{\left( 1-u\right) e^{xt}}{e^{t}-u}=\sum_{n=0}^{\infty }H_{n}(x;u)\frac
t^{n}}{n!}, \label{mt2}
\end{equation
with, of course, $H_{n}(0;u)=H_{n}(u)$ denotes the so-called Eulerian
numbers cf. (\cite{Carlitz}, \cite{Carlitz1952}, \cite{Carlitz1953G}, \cit
{Carlitz1976}, \cite{KimmskimlcjangJIA}, \cite{YsimsekKim}, \cit
{KimSimskJKM}, \cite{SimsekBKMS}, \cite{SimsekJNT}, \cite{srivas18}, \cit
{Tsumura}). Substituting $u=-1$, into (\ref{mt2}), we hav
\begin{equation*}
H_{n}(x;-1)=E_{n}(x)
\end{equation*
where $E_{n}(x)$\ denotes Euler polynomials which are defined by means of
the following generating function
\begin{equation}
\frac{2e^{xt}}{e^{t}+1}=\sum_{n=0}^{\infty }E_{n}(x)\frac{t^{n}}{n!}
\label{1Ssf}
\end{equation
where $\left\vert t\right\vert <\pi $\ cf. \cite{AgohDilcher}-\cite{walum}.
\end{remark}
The following elementary properties of the generalized Eulerian type
polynomials and numbers are derived from their generating functions in (\re
{4ge1}).
\begin{theorem}
(\textit{Recurrence relation} for the generalized Eulerian type numbers):
For $n=0$, we hav
\begin{equation*}
\mathcal{H}_{0}(u;a,b;\lambda )=\left\{
\begin{array}{c}
\frac{1-u}{\lambda -u}\text{ if }a=1, \\
\\
\frac{u}{\lambda -u}\text{ if }a\neq 1
\end{array
\right.
\end{equation*
For $n>0$, following the usual convention of symbolically replacing $\left(
\mathcal{H}(u;a,b;\lambda )\right) ^{n}$ by $\mathcal{H}_{n}(u;a,b;\lambda )
, we hav
\begin{equation*}
\lambda \left( \ln b+\mathcal{H}(u;a,b;\lambda )\right) ^{n}-u\mathcal{H
_{n}(u;a,b;\lambda )=\left( \ln a\right) ^{n}.
\end{equation*}
\end{theorem}
\begin{proof}
By using (\ref{4ge1}), we obtai
\begin{equation*}
\sum_{n=0}^{\infty }\left( \ln a\right) ^{n}\frac{t^{n}}{n!
-u=\sum_{n=0}^{\infty }\left( \lambda \left( \ln b+\mathcal{H}(u;a,b;\lambda
)\right) ^{n}-u\mathcal{H}_{n}(u;a,b;\lambda )\right) \frac{t^{n}}{n!}.
\end{equation*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
By differentiating both sides of equation (\ref{4ge1}) with respect to the
variable $x$, we obtain the following higher order differential equation
\begin{equation}
\frac{\partial ^{j}}{\partial x^{j}}F_{\lambda }(t,x;u,a,b,c)=\left( \ln
\left( c^{t}\right) \right) ^{j}F_{\lambda }(t,x;u,a,b,c). \label{F1}
\end{equation
From this equation, we arrive at higher order derivative of \ the
generalized Eulerian type polynomials by the following theorem:
\begin{theorem}
\label{TEo3} Let $n$, $j\in \mathbb{N}$ with $j\leq n$. Then we hav
\begin{equation*}
\frac{\partial ^{j}}{\partial x^{j}}\mathcal{H}_{n}(x;u;a,b,c;\lambda
)=\left\{ n\right\} _{j}\left( \ln \left( c\right) \right) ^{j}\mathcal{H
_{n-j}(x;u;a,b,c;\lambda ).
\end{equation*}
\end{theorem}
\begin{proof}
Combining (\ref{4ge1}) and (\ref{F1}), we hav
\begin{equation*}
\sum_{n=0}^{\infty }\frac{\partial ^{j}}{\partial x^{j}}\mathcal{H
_{n}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!}=\left( \ln c\right)
^{j}\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,c;\lambda )\frac{t^{n+j}}{n!
.
\end{equation*
From the above equation, we ge
\begin{equation*}
\sum_{n=0}^{\infty }\frac{\partial ^{j}}{\partial x^{j}}\mathcal{H
_{n}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!}=\left( \ln c\right)
^{j}\sum_{n=0}^{\infty }\left\{ n\right\} _{j}\mathcal{H}_{n-j}(x;u;a,b,c
\lambda )\frac{t^{n}}{n!}.
\end{equation*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
Setting $j=1$ in Theorem \ref{TEo3}, we hav
\begin{equation*}
\frac{\partial }{\partial x}\mathcal{H}_{n}(x;u;a,b,c;\lambda )=n\mathcal{H
_{n-1}(x;u;a,b,c;\lambda )\ln \left( c\right) .
\end{equation*
In their special case whe
\begin{equation*}
a=\lambda =1\text{ and }b=c=e,
\end{equation*
Theorem \ref{TEo3}\ is reduced to the following well known result
\begin{equation*}
\frac{\partial ^{j}}{\partial x^{j}}H_{n}(x;u)=\frac{n!}{(n-j)!}H_{n-j}(x;u)
\end{equation*
cf. \cite[Eq-(3.5)]{Carlitz}. Substituting $j=1$ into the above equation, we
hav
\begin{equation*}
\frac{\partial }{\partial x}H_{n}(x;u)=nH_{n-1}(x;u)
\end{equation*
cf. (\cite[Eq-(3.5)]{Carlitz}, \cite{burakSimsek}).
\end{remark}
\begin{theorem}
\label{t8} The following explicit representation formula holds true
\begin{eqnarray*}
&&\left( x\ln c+\ln a\right) ^{n}-ux^{n}\left( \ln c\right) ^{n} \\
&=&\lambda \left( x\ln c+\ln b+\mathcal{H}(u;a,b;\lambda )\right)
^{n}-u\left( x\ln c+\mathcal{H}(u;a,b;\lambda )\right) ^{n}.
\end{eqnarray*}
\end{theorem}
\begin{proof}
By using (\ref{4ge1}) and the \textit{umbral calculus convention}, we obtai
\begin{equation*}
\frac{a^{t}-u}{\lambda b^{t}-u}=e^{H\left( u;a,b;\lambda \right) t}.
\end{equation*
From the above equation, we ge
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\left( \left( \ln a+x\ln c\right) ^{n}-u\left( x\ln
c\right) \right) \frac{t^{n}}{n!} \\
&=&\sum_{n=0}^{\infty }\left( \lambda \left( \mathcal{H}\left( u;a,b;\lambda
\right) +\ln b+x\ln c\right) ^{n}-u\left( \mathcal{H}_{n}\left(
u;a,b;\lambda \right) +x\ln c\right) ^{n}\right) \frac{t^{n}}{n!}.
\end{eqnarray*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
By substituting $a=\lambda =1$ and $b=c=e$ into Theorem \ref{t8}, we hav
\begin{equation}
\left( 1-u\right) x^{n}=H_{n}(x+1;u)-uH_{n}(x;u) \label{at8}
\end{equation
cf. (\cite[Eq-(3.3)]{Carlitz}, \cite{Tsumura}). By setting $u=-1$ in the
above equation, we hav
\begin{equation*}
2x^{n}=E_{n}(x+1)+E_{n}(x)
\end{equation*
a result due to Shiratani \cite{K. Shiratani}. By using (\ref{at8}), Carlitz
\cite{Carlitz} studied on the \textbf{Mirimonoff polynomial} $f_{n}(0,m)$
which is defined b
\begin{eqnarray*}
f_{n}(x,m) &=&\dsum\limits_{j=0}^{m-1}(x+j)^{n}u^{m-j-1} \\
&=&\frac{H_{n}(x+m;u)-u^{m}H_{n}(x;u)}{1-u}.
\end{eqnarray*
By applying Theorem \ref{t8}, one may generalize the Mirimonoff polynomial.
\end{remark}
\begin{theorem}
The following explicit representation formula holds true
\begin{equation}
\mathcal{H}_{n}(x;u;a,b,c;\lambda )=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( x\ln c\right) ^{n-j}\mathcal{H}_{j}(u;a,b,c;\lambda ).
\label{Te}
\end{equation}
\end{theorem}
\begin{proof}
By using (\ref{4ge1}), we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{H}_{n}(u;a,b,c;\lambda )\frac{t^{n}}{n!
\sum_{n=0}^{\infty }\left( \ln c\right) ^{n}\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!}.
\end{equation*
From the above equation, we obtai
\begin{equation*}
\sum_{n=0}^{\infty }\left( \sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( x\ln c\right) ^{n-j}\mathcal{H}_{j}(u;a,b,c;\lambda )\right)
\frac{t^{n}}{n!}=\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,c;\lambda
\frac{t^{n}}{n!}.
\end{equation*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
Substituting $a=\lambda =1$ and $b=c=e$ into (\ref{Te}), we hav
\begin{equation*}
H_{n}(x;u)=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) x^{n-j}H_{j}(u)
\end{equation*
cf. (\cite{Carlitz}, \cite{Carlitz1952}, \cite{Carlitz1953G}, \cit
{Carlitz1976}, \cite{KimmskimlcjangJIA}, \cite{YsimsekKim}, \cit
{KimSimskJKM}, \cite{burakSimsek}, \cite{SimsekBKMS}, \cite{SimsekJNT}, \cit
{srivas18}, \cite{Tsumura}).
\end{remark}
\begin{remark}
From (\ref{Te}), we easily ge
\begin{equation*}
\mathcal{H}_{n}(x;u;a,b,c;\lambda )=\left( \mathcal{H}(u;a,b,c;\lambda
)+x\ln c\right) ^{n},
\end{equation*
where after expansion of the right member, $\mathcal{H}^{n}(u;a,b,c;\lambda
) $ is replaced by $\mathcal{H}_{n}(u;a,b,c;\lambda )$, we use this
convention frequently throughout of this paper.
\end{remark}
\begin{theorem}
\begin{equation}
\mathcal{H}_{n}(x+y;u;a,b,c;\lambda )=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( y\ln c\right) ^{n-j}\mathcal{H}_{j}(x;u;a,b,c;\lambda ).
\label{1Ssa}
\end{equation}
\end{theorem}
\begin{proof}
By using (\ref{4ge1}), we hav
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{H}_{n}(x+y;u;a,b,c;\lambda )\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\left( y\ln c\right) ^{n}\frac{t^{n}}{n!
.\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!}.
\end{equation*
Therefor
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{H}_{n}(x+y;u;a,b,c;\lambda )\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( y\ln c\right) ^{n-j}\mathcal{H}_{j}(x,u;a,b,c;\lambda )\frac
t^{n}}{n!}.
\end{equation*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
In the special case when $a=\lambda =1$ and $b=c=e$, equation (\ref{1Ssa})\
is reduced to the following result
\begin{equation*}
H_{n}(x+y)=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) y^{n-j}H_{j}(x,u)
\end{equation*
cf. \cite[Eq-(3.6)]{Carlitz}. Substituting $u=-1$ into the above equation,
we get the following well-known result
\begin{equation}
E_{n}(x+y)=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) y^{n-j}E_{j}(x). \label{4Eq}
\end{equation}
\end{remark}
By using (\ref{4ge1}), we define the following functional equation
\begin{equation}
F_{\lambda ^{2}}(t,x;u^{2},a^{2},b^{2},c)c^{yt}=F_{\lambda
}(t,x;u,a,b,c)F_{\lambda }(t,y;-u,a,b,c). \label{4EqY}
\end{equation}
\begin{theorem}
\begin{equation}
\mathcal{H}_{n}(x+y;u^{2};a,b,c;\lambda ^{2})=\left( \mathcal{H
(x;u;a,b,c;\lambda )+\mathcal{H}(y;-u;a,b,c;\lambda )\right) ^{n}.
\label{4Eqy1}
\end{equation}
\end{theorem}
\begin{proof}
Combining (\ref{4EqY}) and (\ref{1Ssa}), we easily arrive at the desired
result.
\end{proof}
\begin{remark}
In the special case when $a=\lambda =1$ and $b=c=e$, equation (\ref{4Eqy1})\
is reduced to the following result
\begin{equation*}
H_{n}(x+y;u^{2})=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) H_{j}(x;u)H_{n-j}(y;-u)
\end{equation*
cf. \cite[Eq-(3.17)]{Carlitz}.
\end{remark}
\begin{theorem}
\label{TeoE
\begin{equation*}
(-1)^{n}\mathcal{H}_{n}(1-x;u^{-1};a,b,c;\lambda ^{-1})=\lambda
\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( \ln \left( \frac{b}{a}\right) \right) ^{n-j}\mathcal{H
_{j}(x-1,u;a,b,c;\lambda ).
\end{equation*}
\end{theorem}
\begin{proof}
By using (\ref{4ge1}), we obtai
\begin{equation*}
\frac{\left( a^{-t}-u^{-1}\right) c^{-\left( 1-x\right) t}}{\lambda
^{-1}b^{-t}-u^{-1}}=\lambda \left( \frac{b}{a}\right) ^{t}\sum_{n=0}^{\infty
}\mathcal{H}_{n}(x-1;u;a,b,c;\lambda )\frac{t^{n}}{n!}.
\end{equation*
From the above equation, we ge
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\mathcal{H}_{n}(1-x;u^{-1};a,b,c;\lambda ^{-1})\frac
(-1)^{n}t^{n}}{n!} \\
&=&\lambda \left( \sum_{n=0}^{\infty }\mathcal{H}_{n}(x-1;u;a,b,c;\lambda
\frac{t^{n}}{n!}\right) \left( \sum_{n=0}^{\infty }\left( \ln \left( \frac{
}{a}\right) \right) ^{n}\frac{t^{n}}{n!}\right) .
\end{eqnarray*
Therefor
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }(-1)^{n}\mathcal{H}_{n}(1-x;u^{-1};a,b,c;\lambda ^{-1}
\frac{t^{n}}{n!} \\
&=&\sum_{n=0}^{\infty }\left( \lambda \sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( \ln \left( \frac{b}{a}\right) \right) ^{n-j}\mathcal{H
_{j}(x-1,u;a,b,c;\lambda )\right) \frac{t^{n}}{n!}.
\end{eqnarray*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
In their special case when $a=\lambda =1$ and $b=c=e$, Theorem \ref{TeoE}\
is reduced to the following result
\begin{equation*}
(-1)^{n}H_{n}(1-x;u^{-1})=H_{n}(x-1,u)
\end{equation*
cf. \cite[Eq-(3.7)]{Carlitz}. Substituting $u=-1$ into the above equation,
we get the following well-known result
\begin{equation*}
(-1)^{n}E_{n}(1-x)=E_{n}(x)
\end{equation*
cf. (\cite[Eq-(3.7)]{Carlitz}, \cite{http}, \cite{Parashar}, \cite{K.
Shiratani}, \cite{Srivastava2011}).
\end{remark}
\begin{theorem}
\label{TeoE-1
\begin{equation*}
\mathcal{H}_{n}\left( \frac{x+y}{2};u^{2};a,b,c;\lambda ^{2}\right)
=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \frac{\mathcal{H}_{j}(x;u;a,b,c;\lambda )\mathcal{H
_{n-j}(y;-u;a,b,c;\lambda )}{2^{n}}.
\end{equation*}
\end{theorem}
\begin{proof}
By using (\ref{4ge1}), we ge
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\mathcal{H}_{n}\left( \frac{x+y}{2};u^{2};a,b,c
\lambda ^{2}\right) \frac{2^{n}t^{n}}{n!} \\
&=&\sum_{n=0}^{\infty }\left( \sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{H}_{j}(x;u;a,b,c;\lambda )\mathcal{H}_{n-j}(y;-u;a,b,c
\right) \frac{t^{n}}{n!}.
\end{eqnarray*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
When $a=\lambda =1$ and $b=c=e$, Theorem \ref{TeoE-1}\ is reduced to the
following result
\begin{equation*}
\mathcal{H}_{n}\left( \frac{x+y}{2};u^{2}\right) =2^{-n}\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{H}_{j}(x;u)\mathcal{H}_{n-j}(y;-u),
\end{equation*
cf. \cite[Eq-(3.17)]{Carlitz}.
\end{remark}
\subsection{Multiplication formulas for normalized polynomials}
In this section, using generating functions, we derive \textit
multiplication formulas} in terms of the normalized polynomials which are
related to the generalized Eulerian type polynomials, the Bernoulli and the
Euler polynomials.
\begin{theorem}
\label{T11}(Multiplication formula) Let $y\in \mathbb{N}$. Then we hav
\begin{eqnarray}
&&\mathcal{H}_{n}(yx;u;a,b,b;\lambda ) \label{mMF} \\
&=&y^{n}\dsum\limits_{k=0}^{n}\dsum\limits_{j=0}^{y-1}\left(
\begin{array}{c}
n \\
\end{array
\right) \frac{\lambda ^{j}\left( \ln a\right) ^{n-k}}{u^{j+1-y}-u^{j+1}
\mathcal{H}_{k}\left( x+\frac{j}{y};u^{y};a,b,b;\lambda ^{y}\right) \notag
\\
&&\times \left( H_{n-k}\left( \frac{1}{y};u^{y}\right) -uH_{n-k}\left(
u^{y}\right) \right) , \notag
\end{eqnarray
where $H_{n}\left( x;u\right) $ and $H_{n}\left( u\right) $ denote the
Eulerian polynomials and numbers, respectively.
\end{theorem}
\begin{proof}
Substituting $c=b$ into (\ref{4ge1}), we hav
\begin{equation}
\dsum\limits_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n}}
n!}=\frac{\left( a^{t}-u\right) b^{xt}}{\lambda b^{t}-u}=\left( \frac{a^{t}-
}{-u}\right) \frac{\left( a^{t}-u\right) b^{xt}}{1-\frac{\lambda b^{t}}{u}}.
\label{MT1}
\end{equation
By using the following finite geometric serie
\begin{equation*}
\dsum\limits_{j=0}^{y-1}\left( \frac{\lambda b^{t}}{u}\right) ^{j}=\frac
1-\left( \frac{\lambda b^{t}}{u}\right) ^{y}}{1-\frac{\lambda b^{t}}{u}},
\end{equation*
on the right-hand side of (\ref{MT1}), we obtai
\begin{equation*}
\dsum\limits_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n}}
n!}=\frac{\left( a^{t}-u\right) b^{xt}}{-u\left( 1-\left( \frac{\lambda b^{t
}{u}\right) ^{y}\right) }\dsum\limits_{j=0}^{y-1}\left( \frac{\lambda b^{t}}
u}\right) ^{j}.
\end{equation*
From this equation, we ge
\begin{equation*}
\dsum\limits_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n}}
n!}=\frac{\left( a^{t}-u\right) }{\left( a^{yt}-u^{y}\right)
\dsum\limits_{j=0}^{y-1}\frac{\lambda ^{j}}{u^{j+1-y}}\frac{\left(
a^{yt}-u^{y}\right) b^{yt\left( \frac{x+j}{y}\right) }}{\left( \lambda
b^{yt}-u^{y}\right) }.
\end{equation*
Now by making use of the generating functions (\ref{4ge1}) and (\ref{mt2})
on the right-hand side of the above equation, we obtai
\begin{eqnarray*}
&&\dsum\limits_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n
}{n!} \\
&=&\frac{1}{1-u^{y}}\dsum\limits_{j=0}^{y-1}\frac{\lambda ^{j}}{u^{j+1-y}
\left( \dsum\limits_{n=0}^{\infty }\mathcal{H}_{n}\left( \frac{x+j}{y
;u^{y};a,b,b;\lambda ^{y}\right) \frac{y^{n}t^{n}}{n!}\right) \\
&&\times \left( \dsum\limits_{n=0}^{\infty }\left( H_{n}\left( \frac{1}{y
;u^{y}\right) -uH_{n}\left( u^{y}\right) \right) \frac{\left( y\ln a\right)
^{n}t^{n}}{n!}\right) .
\end{eqnarray*
Therefor
\begin{eqnarray*}
&&\dsum\limits_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n
}{n!} \\
&=&\dsum\limits_{n=0}^{\infty
}\dsum\limits_{k=0}^{n}\dsum\limits_{j=0}^{y-1}\left(
\begin{array}{c}
n \\
\end{array
\right) \frac{y^{n}\lambda ^{j}\left( \ln a\right) ^{n-k}}{u^{j+1-y}-u^{j+1}
\mathcal{H}_{k}\left( \frac{x+j}{y};u^{y};a,b,b;\lambda ^{y}\right) \\
&&\times \left( H_{n-k}\left( \frac{1}{y};u^{y}\right) -uH_{n-k}\left(
u^{y}\right) \right) \frac{t^{n}}{n!}.
\end{eqnarray*
By equating the coefficients of $\frac{t^{n}}{n!}$\ on both sides, we ge
\begin{eqnarray*}
&&\mathcal{H}_{n}(x;u;a,b,b;\lambda ) \\
&=&\dsum\limits_{k=0}^{n}\dsum\limits_{j=0}^{y-1}\left(
\begin{array}{c}
n \\
\end{array
\right) \frac{y^{n}\lambda ^{j}\left( \ln a\right) ^{n-k}}{u^{j+1-y}-u^{j+1}
\mathcal{H}_{k}\left( \frac{x+j}{y};u^{y};a,b,b;\lambda ^{y}\right) \\
&&\times \left( H_{n-k}\left( \frac{1}{y};u^{y}\right) -uH_{n-k}\left(
u^{y}\right) \right) .
\end{eqnarray*
Finally, by replacing $x$ by $yx$ on both sides of the above equation, we
arrive at the desired result.
\end{proof}
\begin{remark}
By substituting $a=1$ into Theorem \ref{T11}, for $n=k$, we obtai
\begin{equation}
\mathcal{H}_{n}(yx;u;1,b,b;\lambda )=y^{n}u^{y-1}\frac{1-u}{1-u^{y}
\dsum\limits_{j=0}^{y-1}\frac{\lambda ^{j}}{u^{j}}\mathcal{H}_{n}\left( x
\frac{j}{y};u^{y};1,b,b;\lambda ^{y}\right) . \label{mmF}
\end{equation
By substituting $b=e$ and $\lambda =1$ into the above equation, we arrive at
the multiplication formula for the Eulerian polynomial
\begin{equation}
H_{n}(yx;u)=y^{n}u^{y-1}\frac{\left( 1-u\right) }{1-u^{y}
\dsum\limits_{j=0}^{y-1}\frac{1}{u^{j}}H_{n}\left( x+\frac{j}{y
;u^{y}\right) , \label{MF-0}
\end{equation
cf. (\cite{Carlitz1953G}, \cite[Eq-(3.12)]{Carlitz}). If $u=-1$, then the
above equation reduces to the well known multiplication formula for the
Euler polynomials: for $y$ is an odd positive integer, we hav
\begin{equation}
E_{n}(yx)=y^{n}\dsum\limits_{j=0}^{y-1}(-1)^{j}E_{n}\left( x+\frac{j}{y
\right) , \label{MF}
\end{equation
where $E_{n}(x)$ denotes the Euler polynomials in the usual notation. If$\ y$
is an even positive integer, we hav
\begin{equation}
E_{n}(yx)=\frac{2y^{n-1}}{n}\dsum\limits_{j=0}^{y-1}(-1)^{j}B_{n}\left( x
\frac{j}{y}\right) , \label{MF2}
\end{equation
where $B_{n}(x)$ and $E_{n}(x)$ denote the Bernoulli polynomials and Euler
polynomials, respectively, cf. (\cite{Carlitz1952}, \cit
{SrivastavaKurtSimsek}).
\end{remark}
To prove the multiplication formula of the generalized Apostol Bernoulli
polynomials, we need the following generating function which is defined by
Srivastava et al. \cite[pp. 254, Eq. (20)]{SrivastawaGargeSC}:
\begin{definition}
\label{DefBER}Let $a,b,c\in \mathbb{R}^{+}$ with $a\neq b,$ $x\in \mathbb{R}$
and $n\in \mathbb{N}_{0}$. Then the generalized Bernoulli polynomials
\mathfrak{B}_{n}^{(\alpha )}(x;\lambda ;a,b,c)$ of order $\alpha \in \mathbb
C}$ are defined by means of the following generating functions
\begin{equation}
f_{B}(x,a,b,c;\lambda ;\alpha )=\left( \frac{t}{\lambda b^{t}-a^{t}}\right)
^{\alpha }c^{xt}=\sum_{n=0}^{\infty }\mathfrak{B}_{n}^{(\alpha )}(x;\lambda
;a,b,c)\frac{t^{n}}{n!}, \label{1S}
\end{equation
wher
\begin{equation*}
\left\vert t\ln (\frac{a}{b})+\ln \lambda \right\vert <2\pi
\end{equation*
an
\begin{equation*}
1^{\alpha }=1.
\end{equation*}
\end{definition}
Observe that if we set $\lambda =1$ in (\ref{1S}), we hav
\begin{equation}
\left( \frac{t}{b^{t}-a^{t}}\right) ^{\alpha }c^{xt}=\sum_{n=0}^{\infty
\mathfrak{B}_{n}^{(\alpha )}(x;a,b,c)\frac{t^{n}}{n!}. \label{9}
\end{equation
If we set $x=0$ in (\ref{9}), we obtai
\begin{equation}
\left( \frac{t}{b^{t}-a^{t}}\right) ^{\alpha }=\sum_{n=0}^{\infty }\mathfrak
B}_{n}^{(\alpha )}(a,b)\frac{t^{n}}{n!}, \label{8}
\end{equation
with of course, $\mathfrak{B}_{n}^{(\alpha )}(x;a,b,c)=\mathfrak{B
_{n}^{(\alpha )}(a,b)$, cf. (\cite{luo14}-\cite{lou15}, \cite{KimSimskJKM},
\cite{YsimsekKim}, \cite{kurtSimsek}, \cite{Mali}, \cite{OzdenAML}, \cit
{OzdenSrivastava}, \cite{srivas18}, \cite{srivastava11}, \cit
{Srivastava2011}, \cite{SrivastawaGargeSC}). If we set $\alpha =1$ in (\re
{8}) and (\ref{9}), we hav
\begin{equation}
\frac{t}{b^{t}-a^{t}}=\sum_{n=0}^{\infty }\mathfrak{B}_{n}(a,b)\frac{t^{n}}
n!} \label{4}
\end{equation
and
\begin{equation}
\left( \frac{t}{b^{t}-a^{t}}\right) c^{xt}=\sum_{n=0}^{\infty }\mathfrak{B
_{n}(x;a,b,c)\frac{t^{n}}{n!}, \label{5}
\end{equation
which have been studied by Luo et al. \cite{luo14}-\cite{lou15}. Moreover,
by substituting $a=1$ and $b=c=e$ into (\ref{1S}), then we arrive at the
Apostol-Bernoulli polynomials $\mathcal{B}_{n}(x;\lambda )$, which are
defined by means of the following generating functio
\begin{equation*}
\left( \frac{t}{\lambda e^{t}-1}\right) e^{xt}=\sum_{n=0}^{\infty }\mathcal{
}_{n}(x;\lambda )\frac{t^{n}}{n!},
\end{equation*
These polynomials $\mathcal{B}_{n}(x;\lambda )$ have been introduced and
investigated by many Mathematicians cf. (\cite{apostol}, \cite{KimChiARXIV},
\cite{KimkimJang}, \cite{KimSimskJKM}, \cite{burakSimsek}, \cite{luo13},
\cite{OzdenSrivastava}, \cite{SimsekJMAA}, \cite{srivas18}). When $a=\lambda
=1$ and $b=c=e$ into (\ref{4}) and (\ref{5}), $\mathfrak{B}_{n}(a,b)$ and
\mathfrak{B}_{n}(x;a,b,c)$ reduce to the classical Bernoulli numbers and the
classical Bernoulli polynomials, respectively, cf. \cite{AgohDilcher}-\cit
{walum}.
\begin{remark}
The constraints on $\left\vert t\right\vert $, which we have used in
Definition \ref{DefBER} and (\ref{1Ssf}), are meant to ensure that the
generating function in (\ref{9})and (\ref{1Ssf}) are analytic throughout the
prescribed open disks in complex $t$-plane (centred at the origin $t=0$) in
order to have the corresponding convergent Taylor-Maclaurin series expansion
(about the origin $t=0$) occurring on the their right-hand side (with a
positive radius of convergence) cf. \cite{srivastava11}.
\end{remark}
\begin{theorem}
\label{T11a}Let $y\in \mathbb{N}$. Then we hav
\begin{equation*}
\mathfrak{B}_{n}(yx;\lambda
;a,b,b)=\dsum\limits_{l=0}^{n}\dsum\limits_{j=0}^{y-1}\left(
\begin{array}{c}
n \\
\end{array
\right) \lambda ^{j}y^{l-1}\left( (y-1-j)\ln a\right) ^{n-l}\mathfrak{B
_{l}\left( x+\frac{j}{y};\lambda ^{y};a,b,b\right) .
\end{equation*}
\end{theorem}
\begin{proof}
Substituting $c=b$ and $\alpha =1$ into (\ref{1S}), we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathfrak{B}_{n}(x;\lambda ;a,b,c)\frac{t^{n}}{n!}=\frac
1}{y}\sum_{j=0}^{y-1}\lambda ^{j}\frac{yt}{\lambda ^{y}b^{yt}-a^{yt}
b^{\left( \frac{x+j}{y}\right) yt}a^{t(y-j-1)}.
\end{equation*
Therefor
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\mathfrak{B}_{n}(x;\lambda ;a,b,c)\frac{t^{n}}{n!} \\
&=&\sum_{n=0}^{\infty }\dsum\limits_{l=0}^{n}\dsum\limits_{j=0}^{y-1}\left(
\begin{array}{c}
n \\
\end{array
\right) \lambda ^{j}\left( (y-1-j)\ln a\right) ^{n-l}y^{l-1}\mathfrak{B
_{l}\left( \frac{x+j}{y};\lambda ^{y};a,b,b\right) \frac{t^{n}}{n!}.
\end{eqnarray*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we ge
\begin{equation*}
\mathfrak{B}_{n}^{(\alpha )}(x;\lambda
;a,b,c)=\dsum\limits_{l=0}^{n}\dsum\limits_{j=0}^{y-1}\left(
\begin{array}{c}
n \\
\end{array
\right) \lambda ^{j}\left( (k-1-j)\ln a\right) ^{n-l}y^{l-1}\mathfrak{B
_{l}\left( \frac{x+j}{k};\lambda ^{y};a,b,b\right) .
\end{equation*
By replacing $x$ by $yx$ on both sides of the above equation, we arrive at
the desired result.
\end{proof}
\begin{remark}
Kurt and Simsek \cite{kurtSimsek} proved multiplication formula for the
generalized Bernoulli polynomials of order $\alpha $. When $a=\lambda =1$
and $b=c=e$ into Theorem \ref{T11a}, we have the multiplication formula for
the Bernoulli polynomials given b
\begin{equation}
B_{n}(yx)=y^{n-1}\dsum\limits_{j=0}^{y-1}B_{n}\left( x+\frac{j}{y}\right) ,
\label{MF1}
\end{equation
cf. (\cite{apostol}, \cite{Carlitz1952}, \cite{Carlitz}, \cite{DERE}, \cit
{KimSimskJKM}, \cite{lou15}, \cite{luo13}, \cite{luo2003}, \cite{Mali}, \cit
{LuoSrivatava2010}, \cite{srivas18}, \cite{SrivastawaGargeSC}).
\end{remark}
If $f$ is a \textit{normalized} polynomial such that it satisfies the formul
\begin{equation}
f_{n}(yx)=y^{n-1}\dsum\limits_{j=0}^{y-1}f_{n}\left( x+\frac{j}{y}\right) ,
\label{w}
\end{equation
then $f$ is the $y$th degree Bernoulli polynomial due to (\ref{MF1}) cf.
\cite{Carlitz1952}, \cite{walum}). According to Nielsen \cite{Carlitz1952},
if a normalized polynomial satisfies (\ref{MF1}) for a single value of $y>1
, then it is identical with $B_{m}(x)$. Consequently, if a normalized
polynomial satisfies (\ref{mmF}) for a single value of $y>1$, then it is
identical with $\mathcal{H}_{n}(x;u;1,b,b;\lambda )$. The formula (\ref{MF2
) is different. Therefore, for $y$ is an even positive integer, Carlitz \cit
[Eq-(1.4)]{Carlitz1952} considered the following equation
\begin{equation*}
g_{n-1}(yx)=-\frac{2y^{n-1}}{n}\dsum\limits_{j=0}^{y-1}(-1)^{j}f_{n}\left( x
\frac{j}{y}\right) ,
\end{equation*
where $g_{n-1}(x)$\ and $f_{n}(x)$ denote the normalized polynomials of
degree $n-1$ and $n$, respectively. More precisely, as Carlitz has pointed
out \cite[p. 184]{Carlitz1952}, if $y$ is a fixed even integer $\geq 2$ and
f_{n}(x)$\ is an arbitrary normalized polynomial of degree $n$, then (\re
{MF2}) determines $g_{n-1}(x)$\ as a normalized polynomial of degree $n-1$.
Thus, for a single value$\ y$, (\ref{MF2}) does not suffice to determine the
normalized polynomials $g_{n-1}(x)$\ and $f_{n}(x)$.
\begin{remark}
According to (\ref{w}), the set of normalized polynomials $\left\{
f_{n}(x)\right\} $ is an Appell set, cf. \cite{Carlitz1952}.
\end{remark}
We now modify (\ref{4ge1}) as follows
\begin{equation}
\frac{\left( a^{t}-\xi \right) c^{xt}}{\lambda b^{t}-\xi
=\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;\xi ;a,b,c;\lambda )\frac{t^{n}}{n!}
\label{M1b}
\end{equation
wher
\begin{equation*}
\xi ^{r}=1,\text{ }\xi \neq 1.
\end{equation*}
The polynomial $\mathcal{H}_{n}(x;\xi ;a,b,c;\lambda )$ is a normalized
polynomial of degree $m$ in $x$. The polynomial \QTR{cal}{H}$_{n}(x;\xi
;1,e,e;1)$ may be called Eulerian polynomials with parameter $\xi $. In
particular we note tha
\begin{equation*}
\mathcal{H}_{n}(x;-1;1,e,e;1)=E_{n}(x)
\end{equation*
since for $a=\lambda =1$, $b=c=e$, equation (\ref{M1b}) reduces to the
generating function for the Euler polynomials.
By means of equation (\ref{mMF}), it is easy to verify the following
multiplication formulas:
If $y$ is an odd positive integer, then we hav
\begin{eqnarray}
\mathcal{H}_{n-1}(yx;\xi ;a,b,b;\lambda ) &=&\frac{y^{n-1}}{n
\dsum\limits_{j=0}^{y-1}\left( \frac{\lambda }{\xi }\right) ^{j}\mathfrak{B
_{n}\left( x+\frac{j}{y};b;\lambda ^{y}\right) \label{M1c} \\
&&-\frac{1}{\xi n}\dsum\limits_{k=0}^{n}\dsum\limits_{j=0}^{y-1}\left( \frac
\lambda }{\xi }\right) ^{j}y^{k-1}(\ln a)^{n-k}\mathfrak{B}_{k}\left( x
\frac{j}{y};b;\lambda ^{y}\right) , \notag
\end{eqnarray
wher
\begin{equation*}
\mathcal{H}_{k}\left( x+\frac{j}{y};\xi ^{y};1,b,b;\lambda ^{y}\right)
\mathfrak{B}_{n}\left( x+\frac{j}{y};b;\lambda ^{y}\right) \text{.}
\end{equation*
If $y$ is an even positive integer, then we hav
\begin{eqnarray}
\mathcal{H}_{n}(yx;\xi ;a,b,b;\lambda ) &=&\frac{y^{n}}{2
\dsum\limits_{j=0}^{y-1}\left( \frac{\lambda }{\xi }\right) ^{j}\mathfrak{E
_{n}\left( x+\frac{j}{y};b;\lambda ^{y}\right) \label{M1d} \\
&&-\frac{1}{2\xi }\dsum\limits_{k=0}^{n}\dsum\limits_{j=0}^{y-1}\left( \frac
\lambda }{\xi }\right) ^{j}y^{k}(\ln a)^{n-k}\mathfrak{E}_{k}\left( x+\frac{
}{y};b;\lambda ^{y}\right) , \notag
\end{eqnarray
wher
\begin{equation*}
\mathcal{H}_{k}\left( x+\frac{j}{y};\xi ^{y};1,b,b;\lambda ^{y}\right)
\mathfrak{E}_{n}\left( x+\frac{j}{y};b;\lambda ^{y}\right) \text{,}
\end{equation*
where $\mathfrak{E}_{n}(x;a,b,c)$ denotes the generalized Euler polynomials,
which are defined by means of the following generating function
\begin{equation*}
\left( \frac{t}{b^{t}-a^{t}}\right) c^{xt}=\sum_{n=0}^{\infty }\mathfrak{E
_{n}(x;a,b,c)\frac{t^{n}}{n!}
\end{equation*
cf. (\cite{luo14}-\cite{lou15}, \cite{Kim Jang}, \cite{kurtSimsek}, \cit
{OzdenSrivastava}, \cite{srivas18}, \cite{srivastava11}, \cit
{Srivastava2011}, \cite{SrivastawaGargeSC}).
\begin{remark}
If we set $a=\lambda =1$ and $b=e$, then (\ref{M1c}) and (\ref{M1d}) reduce
to the following multiplication formulas, respectively
\begin{equation*}
H_{n-1}(yx;\xi )=\frac{y^{n-1}}{n}\left( 1-\frac{1}{\xi }\right)
\dsum\limits_{j=0}^{y-1}\frac{1}{\xi ^{j}}B_{n}\left( x+\frac{j}{y}\right)
\end{equation*
cf. \cite[Eq. (3.3)]{Carlitz1952} an
\begin{equation*}
H_{n}(yx;\xi )=\frac{y^{n}}{2}\left( 1-\frac{1}{\xi }\right)
\dsum\limits_{j=0}^{y-1}\frac{1}{\xi ^{j}}E_{n}\left( x+\frac{j}{y}\right) .
\end{equation*
Let $f_{n}(x)$ and $g_{n}(x)$ be normalized polynomials in the usual way.
Carlitz \cite[Eq. (3.4)]{Carlitz1952} defined the following equation
\begin{equation*}
g_{n-1}(yx)=\frac{(1-\rho )y^{n-1}}{n}\dsum\limits_{j=0}^{y-1}\rho
^{j}f_{n}\left( x+\frac{j}{y}\right) ,
\end{equation*
where $\rho $ is a fixed primitive $r$th root of unity, $r>1$, $y\equiv 0
\func{mod}r)$.
\end{remark}
\begin{remark}
If we set $a=\lambda =1$, $b=c=e$ and $\xi =-1$, then (\ref{M1c}) and (\re
{M1d}) reduce to (\ref{MF2}) and (\ref{MF}).
\end{remark}
\begin{remark}
Walum \cite{walum} defined multiplication formula for periodic functions as
follows
\begin{equation}
\vartheta (y)f(yx)=\dsum\limits_{j(y)}f\left( x+\frac{j}{y}\right) ,
\label{w1}
\end{equation
where $f$ is periodic with period $1$ and $j(y)$ under the summation sign
indicates that $j$ runs through a complete system of residues $\func{mod}y$.
Formulas (\ref{w}), (\ref{w1}) and other multiplication formulas related to
periodic functions and normalized polynomials occur in Franel's formula, in
the theory of the Dedekind sums and Hardy-Berndt sums, in the theory of the
zeta functions and $L$-functions and in the theory of periodic bounded
variation, cf. (\cite{Berndt}, \cite{berndt2}, \cite{walum}).
\end{remark}
\subsection{Generalized Eulerian type numbers and polynomials attached to
Dirichlet character}
In this section, we construct generating function, related to nonnegative
real parameters, for the generalized Eulerian type numbers and polynomials
attached to Dirichlet character. We also give some properties of these
polynomials and numbers.
\begin{definition}
Let $\chi $ be the Dirichlet character of conductor $f\in \mathbb{N}$. Let
x\in \mathbb{R}$, $a,b\in \mathbb{R}^{+},$ $(a\neq b),$ $\lambda \in \mathbb
C}$ and $u\in \mathbb{C\diagdown }\left\{ 1\right\} $. The\ generalized
Eulerian type polynomials $\mathcal{H}_{n,\chi }(x;u;a,b,c;\lambda )$ are
defined by means of the following generating function
\begin{equation}
\mathcal{F}_{\lambda ,\chi }(t,x;u,a,b,c)=\dsum\limits_{j=0}^{f-1}\frac
\left( a^{ft}-u^{f}\right) \chi (j)u^{f-j-1}c^{\left( \frac{x+j}{f}\right)
ft}}{\lambda ^{f}b^{ft}-u^{f}}=\sum_{n=0}^{\infty }\mathcal{H}_{n,\chi
}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!} \label{4ge2}
\end{equation}
with, of cours
\begin{equation*}
\mathcal{H}_{n,\chi }(0;u;a,b,c;\lambda )=\mathcal{H}_{n,\chi
}(u;a,b,c;\lambda ),
\end{equation*
where $\mathcal{H}_{n,\chi }(u;a,b,c;\lambda )$ denotes generalized Eulerian
type numbers.
\end{definition}
\begin{remark}
In the special case when $a=\lambda =1$ and $b=c=e$, the generalized
Eulerian type polynomials $\mathcal{H}_{n,\chi }(x;u;a,b,c;\lambda )$\ are
reduced to the Frobenius Euler polynomials which are defined by means of the
following generating function
\begin{equation*}
\dsum\limits_{j=0}^{f-1}\frac{\left( 1-u^{f}\right) \chi
(j)u^{f-j-1}e^{\left( \frac{x+j}{f}\right) ft}}{e^{ft}-u^{f}
=\sum_{n=0}^{\infty }H_{n,\chi }(x;u)\frac{t^{n}}{n!},
\end{equation*
cf. (\cite{Tsumura}, \cite{KimSimskJKM}, \cite{YsimsekKim}, \cite{SimsekBKMS
, \cite{SimsekJNT}, \cite{srivas18}). Substituting $u=-1$ into the above
equation, we have generating function of the generalized Euler polynomials
attached to Dirichlet character with odd conductor
\begin{equation*}
2\dsum\limits_{j=0}^{f-1}\frac{\chi (j)(-1)^{j}e^{\left( \frac{x+j}{f
\right) ft}}{e^{ft}+1}=\sum_{n=0}^{\infty }E_{n,\chi }(x)\frac{t^{n}}{n!},
\end{equation*
cf. (\cite{Tsumura}, \cite{SimsekBKMS}, \cite{SimsekJNT}, \cite{srivas18}).
\end{remark}
Combining (\ref{4ge1}) and (\ref{4ge2}), we obtain the following functional
equation
\begin{equation*}
\mathcal{F}_{\lambda ,\chi }(t,x;u,a,b,c;)=\dsum\limits_{j=0}^{f-1}\chi
(j)u^{f-j-1}F_{\lambda ^{f}}(ft,\frac{x+j}{f};u^{f},a,b,c).
\end{equation*}
By using the above functional equation we arrive at the following Theorem:
\begin{theorem}
\begin{equation*}
\mathcal{H}_{n,\chi }(x;u;a,b,c;\lambda )=f^{n}\dsum\limits_{j=0}^{f-1}\chi
(j)u^{f-j-1}\mathcal{H}_{n}(\frac{x+j}{f};u^{f};a,b,c;\lambda ^{f}).
\end{equation*}
\end{theorem}
\begin{theorem}
\begin{equation*}
\mathcal{H}_{n,\chi }(x;u;a,b,c;\lambda )=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( x\ln c\right) ^{n-j}\mathcal{H}_{j,\chi }(u;a,b,c;\lambda ).
\end{equation*}
\end{theorem}
\begin{proof}
By using (\ref{4ge2}), we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{H}_{n,\chi }(u;a,b,c;\lambda )\frac{t^{n}}{n!
\sum_{n=0}^{\infty }\left( x\ln c\right) ^{n}\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\mathcal{H}_{n,\chi }(x;u;a,b,c;\lambda )\frac{t^{n}}{n
}.
\end{equation*
From the above equation, we obtai
\begin{equation*}
\sum_{n=0}^{\infty }\left( \sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \left( x\ln c\right) ^{n-j}\mathcal{H}_{j,\chi }(u;a,b,c;\lambda
\right) \frac{t^{n}}{n!}=\sum_{n=0}^{\infty }\mathcal{H}_{n,\chi
}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!}.
\end{equation*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\subsection{\textbf{Recurrence relation for the }generalized Eulerian type
polynomials}
In this section we are going to differentiate (\ref{4ge1}) with respect to
the variable $t$ to derive a recurrence relation for the generalized
Eulerian type polynomials. Therefore, we obtain the following differential
equation
\begin{eqnarray*}
\frac{\partial }{\partial t}F_{\lambda }(t,x;u,a,b,c) &=&\left( \ln a\right)
F_{\lambda }(t,x;u,a,b,c)+\frac{\ln a}{t}f_{B}(x,1,b,c;\frac{\lambda }{u};1)
\\
&&-\frac{\ln \left( b^{\lambda }\right) }{ut}F_{\lambda
}(t,x;u,a,b,c)f_{B}(1,1,b,b;\frac{\lambda }{u};1) \\
&&+\ln \left( c^{x}\right) F_{\lambda }(t,x;u,a,b,c).
\end{eqnarray*
By using this equation, we obtain a recurrence relation for the generalized
Eulerian type polynomials by the following theorem:
\begin{theorem}
Let $n\in \mathbb{N}$. We hav
\begin{eqnarray*}
n\mathcal{H}_{n}(x;u;a,b,c;\lambda ) &=&\left( \ln a\right) \left( n\mathcal
H}_{n-1}(x;u;a,b,c;\lambda )+\mathfrak{B}_{n}(x;\frac{\lambda }{u
;a,b,c)\right) \\
&&-\frac{\lambda \ln b}{u}\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{H}_{j}(x;u;a,b,c;\lambda )\mathfrak{B}_{n-j}(1;\frac
\lambda }{u};1,b,b) \\
&&+\left( \ln \left( c^{nx}\right) \right) \mathcal{H}_{n-1}(x;u;a,b,c
\lambda ),
\end{eqnarray*
where $\mathfrak{B}_{n}(x;\lambda ;a,b,c)$ denotes the generalized Bernoulli
polynomials of order $1$.
\end{theorem}
\begin{remark}
When $a=\lambda =1$ and $b=c=e$, the recurrence relation for the generalized
Eulerian type polynomials is reduced t
\begin{equation*}
nH_{n}(x;u)=nxH_{n-1}(x;u)-\frac{1}{u}\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) H_{j}(x;u)\mathcal{B}_{n-j}(1;\frac{1}{u}).
\end{equation*}
\end{remark}
\section{New identities involving families of polynomials}
In this section, we derive some new identities related to the generalized
Bernoulli polynomials and numbers of order $1$, the Eulerian type
polynomials and the generalized array type polynomials.
\begin{theorem}
\label{T11b} The following relationship holds true
\begin{equation*}
\mathfrak{B}_{n}(x;\lambda ;a,b,b)=\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{H}_{j}(x;\lambda ^{-1};a,\frac{b}{a},\frac{b}{a};1
\mathfrak{B}_{n-j}(x-1;\lambda ;1,a,a).
\end{equation*}
\end{theorem}
\begin{proof}
\begin{equation*}
\sum_{n=0}^{\infty }\mathfrak{B}_{n}(x;\lambda ;a,b,b)\frac{t^{n}}{n!
=\left( \frac{ta^{(x-1)t}}{\lambda a^{t}-1}\right) \left( \frac{\left(
a^{t}-\lambda ^{-1}\right) \left( \frac{b}{a}\right) ^{xt}}{\left( \frac{b}{
}\right) ^{t}-\lambda ^{-1}}\right) .
\end{equation*
Combining (\ref{1S}) and (\ref{4ge1}) with the above equation, we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathfrak{B}_{n}(x;\lambda ;a,b,b)\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\mathfrak{B}_{n}(x-1;\lambda ;1,a,a)\frac{t^{n}}{n!
\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;\lambda ^{-1};a,\frac{b}{a},\frac{b}{a
;1)\frac{t^{n}}{n!}.
\end{equation*
Therefor
\begin{equation*}
\sum_{n=0}^{\infty }\mathfrak{B}_{n}(x;\lambda ;a,b,b)\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\left( \sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \mathcal{H}_{j}(x;\lambda ^{-1};a,\frac{b}{a},\frac{b}{a};1
\mathfrak{B}_{n-j}(x-1;\lambda ;1,a,a)\right) \frac{t^{n}}{n!}.
\end{equation*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
Relationship between the generalized Bernoulli numbers and the Frobenius
Euler numbers is given by the following result:
\begin{theorem}
The following relationship holds true
\begin{equation}
\mathfrak{B}_{n}(\lambda ;a,b)=\frac{1}{\lambda -1}\sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) j\left( \ln a^{-1}\right) ^{n-j}\left( \ln \left( \frac{b}{a}\right)
\right) ^{j}H_{j-1}\left( \lambda ^{-1}\right) . \label{1Ssd}
\end{equation}
\end{theorem}
\begin{proof}
By using (\ref{1S}), we obtai
\begin{equation*}
\sum_{n=0}^{\infty }\mathfrak{B}_{n}(\lambda ;a,b)\frac{t^{n}}{n!}=\frac
ta^{-t}}{\lambda -1}\left( \frac{1-\lambda ^{-1}}{e^{t\ln \left( \frac{b}{a
\right) }-\lambda ^{-1}}\right) .
\end{equation*}
From the above equation, we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathfrak{B}_{n}(\lambda ;a,b)\frac{t^{n}}{n!}=\frac{1}
\lambda -1}\sum_{n=0}^{\infty }\left( \ln \left( \frac{1}{a}\right) \right)
^{n}\frac{t^{n}}{n!}\sum_{n=0}^{\infty }n\mathcal{H}_{n}(\lambda
^{-1})\left( \ln \left( \frac{b}{a}\right) \right) ^{n}\frac{t^{n}}{n!}.
\end{equation*
Therefor
\begin{equation*}
\sum_{n=0}^{\infty }\mathfrak{B}_{n}(\lambda ;a,b)\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\left( \sum_{j=0}^{n}\left(
\begin{array}{c}
n \\
\end{array
\right) \frac{j\left( \ln a^{-1}\right) ^{n-j}\left( \ln \left( \frac{b}{a
\right) \right) ^{j}}{\lambda -1}H_{j-1}\left( \lambda ^{-1}\right) \right)
\frac{t^{n}}{n!}.
\end{equation*}
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
By substituting $a=1$ and $b=e$ into (\ref{1Ssd}), we hav
\begin{equation*}
\mathcal{B}_{n}(\lambda )=\frac{n}{\lambda -1}H_{n-1}(\lambda ^{-1}),
\end{equation*
cf. \cite{KimSimskJKM}.
\end{remark}
Relationship between the generalized Eulerian type polynomials and
generalized array type polynomials are given by the following theorem:
\begin{theorem}
The following relationship holds true
\begin{equation*}
\mathcal{H}_{n}(x;u;a,b,b;\lambda )=\sum_{k=0}^{\infty }\sum_{m=0}^{\infty
}\sum_{d=0}^{n}\left(
\begin{array}{c}
m+k-1 \\
\end{array
\right) \left(
\begin{array}{c}
n \\
\end{array
\right) \frac{k!\left( \ln a^{m}\right) ^{n-d}}{u^{m+k}}\mathcal{S
_{k}^{d}(x;a,b;\lambda ).
\end{equation*}
\end{theorem}
\begin{proof}
From (\ref{4ge1}), we obtai
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,c;\lambda )\frac{t^{n}}{n!
=\sum_{k=0}^{\infty }\left( \frac{\lambda b^{t}-a^{t}}{u-a^{t}}\right)
^{k}b^{xt}.
\end{equation*
Combining (\ref{ab0}) with the above equation, we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n}}{n!
=\sum_{k=0}^{\infty }\frac{k!}{\left( u-a^{t}\right) ^{k}}\sum_{n=0}^{\infty
}\mathcal{S}_{k}^{n}(x;a,b;\lambda )\frac{t^{n}}{n!}.
\end{equation*
From the above equation, we ge
\begin{equation*}
\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n}}{n!
=\sum_{n=0}^{\infty }\sum_{k=0}^{\infty }\frac{k!\mathcal{S
_{k}^{n}(x;a,b;\lambda )}{u^{k}\left( 1-\frac{a^{t}}{u}\right) ^{k}}\frac
t^{n}}{n!}\text{.}
\end{equation*
Now we assume $\left\vert \frac{a^{t}}{u}\right\vert <1$ in the above
equation; thus we ge
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n}}{n!} \\
&=&\sum_{n=0}^{\infty }\sum_{k=0}^{\infty }\sum_{m=0}^{\infty }\left(
\begin{array}{c}
m+k-1 \\
\end{array
\right) \frac{k!\mathcal{S}_{k}^{n}(x;a,b;\lambda )}{u^{k+m}}\frac
a^{mt}t^{n}}{n!}.
\end{eqnarray*
Therefor
\begin{eqnarray*}
&&\sum_{n=0}^{\infty }\mathcal{H}_{n}(x;u;a,b,b;\lambda )\frac{t^{n}}{n!} \\
&=&\sum_{n=0}^{\infty }\left( \sum_{k=0}^{\infty }\sum_{m=0}^{\infty
}\sum_{d=0}^{n}\left(
\begin{array}{c}
m+k-1 \\
\end{array
\right) \left(
\begin{array}{c}
n \\
\end{array
\right) \frac{k!\left( \ln a^{m}\right) ^{n-d}}{u^{m+k}}\mathcal{S
_{k}^{d}(x;a,b;\lambda ).\right) \frac{t^{n}}{n!}.
\end{eqnarray*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
Substituting $a=1$ into the above Theorem and noting that $d=n$, we deduce
the following identity
\begin{equation*}
\mathcal{H}_{n}(x;u;1,b,b;\lambda )=\sum_{k=0}^{\infty }\frac{k!}{(u-1)^{k}
\mathcal{S}_{k}^{n}(x;1,b;\lambda )
\end{equation*
which upon setting $\lambda =1$ and $b=e$, yield
\begin{equation*}
H_{n}(x;u)=\sum_{k=0}^{n}\frac{k!}{(u-1)^{k}}\mathcal{S}_{k}^{n}(x)
\end{equation*
which was found by Chang and Ha \cite[Lemma 1]{Chan}.
\end{remark}
\section{Relationship between the generalized Bernoulli polynomials and the
generalized array type polynomials}
In this section, we give some applications related to the generalized
Bernoulli polynomials, generalized array type polynomials. We derive many
identities involving these polynomials. By using same method with Agoh and
Dilcher's \cite{AgohDilcher}, we give the following Theorem:
\begin{theorem}
\label{L-AD
\begin{equation}
\left( \frac{\lambda b^{t}-a^{t}}{t}\right)
^{k}b^{xt}=\dsum\limits_{n=0}^{\infty }\frac{\mathcal{S}_{k}^{n+k}(x;a,b
\lambda )}{\binom{n+k}{k}}\frac{t^{n}}{n!}. \label{w3}
\end{equation}
\end{theorem}
\begin{proof}
Combining\ (\ref{ab0}) and (\ref{ab1}), we get
\begin{eqnarray*}
\left( \frac{\lambda b^{t}-a^{t}}{t}\right) ^{k}b^{xt} &=&\frac{1}{t^{k}
\dsum\limits_{n=0}^{\infty }\frac{k!}{n!}S_{k}^{n}\left( x,a,b;\lambda
\right) t^{n} \\
&=&\dsum\limits_{n=0}^{\infty }\frac{k!}{n!}S_{k}^{n+k}\left( x,a,b;\lambda
\right) t^{n-k}.
\end{eqnarray*
From the above equation, we arrive at the desired result.
\end{proof}
\begin{remark}
By setting $x=0$, $a=\lambda =1$ and $b=e$, Theorem \ref{L-AD} yields the
corresponding result which is proven by Agoh and Dilcher \cite{AgohDilcher}.
\end{remark}
\begin{theorem}
\label{Tw3
\begin{eqnarray*}
&&\left( n+k\right) \frac{\mathcal{S}_{k}^{n+k}(x;a,b;\lambda )}{\binom{n+k}
k}}-xn\frac{\mathcal{S}_{k}^{n+k-1}(x;a,b;\lambda )}{\binom{n+k-1}{k}} \\
&=&\sum_{j=0}^{n}\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) }{\left(
\begin{array}{c}
j+k-1 \\
k-
\end{array
\right) }\mathcal{S}_{k-1}^{j+k-1}(x;a,b;\lambda )\left( \ln \left(
b^{\lambda k}\right) \left( \ln (b)\right) ^{n-j}-\ln \left( a^{k}\right)
\left( \ln (a)\right) ^{n-j}\right) .
\end{eqnarray*}
\end{theorem}
\begin{proof}
By differentiating both sides of equation (\ref{w3}) with respect to the
variable $t$, after some elementary calculations, we get the formula
asserted by Theorem \ref{Tw3}.
\end{proof}
\begin{theorem}
\label{Teo19} The following relationship holds true
\begin{equation*}
\mathcal{S}_{k-1}^{n}(x+y;a,b;\lambda )=\sum_{j=0}^{n}\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) \left(
\begin{array}{c}
n+k-1 \\
k-
\end{array
\right) }{\left(
\begin{array}{c}
j+k \\
\end{array
\right) }\mathcal{S}_{k}^{j+k}(x;a,b;\lambda )\mathfrak{B}_{n-j}(y;\lambda
;a,b,b).
\end{equation*}
\end{theorem}
\begin{proof}
We se
\begin{equation*}
\left( \frac{\lambda b^{t}-a^{t}}{t}\right) ^{k}b^{xt}\left( \frac{tb^{yt}}
\lambda b^{t}-a^{t}}\right) =\left( \frac{\lambda b^{t}-a^{t}}{t}\right)
^{k-1}b^{(x+y)t}.
\end{equation*
Combining (\ref{w3}) and (\ref{5}) with the above equation, we ge
\begin{equation*}
\sum_{n=0}^{\infty }\frac{\mathcal{S}_{k-1}^{n+k-1}(x+y;a,b;\lambda )}
\binom{n+k-1}{k-1}}\frac{t^{n}}{n!}=\dsum\limits_{n=0}^{\infty }\mathfrak{B
_{n}(y;\lambda ;a,b,b)\frac{t^{n}}{n!}\sum_{n=0}^{\infty }\frac{\mathcal{S
_{k}^{n+k}(x;a,b;\lambda )}{\binom{n+k}{k}}\frac{t^{n}}{n!}.
\end{equation*}
Therefor
\begin{equation*}
\sum_{n=0}^{\infty }\frac{\mathcal{S}_{k-1}^{n+k-1}(x+y;a,b;\lambda )}
\binom{n+k-1}{k-1}}\frac{t^{n}}{n!}=\sum_{n=0}^{\infty }\left( \sum_{j=0}^{n
\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) }{\left(
\begin{array}{c}
j+k \\
\end{array
\right) }\mathcal{S}_{k}^{j+k}(x;a,b;\lambda )\mathfrak{B}_{n-j}(y;\lambda
;a,b,b)\right) \frac{t^{n}}{n!}.
\end{equation*
Comparing the coefficients of $\frac{t^{n}}{n!}$ on both sides of the above
equation, we arrive at the desired result.
\end{proof}
\begin{remark}
By setting $x=y=0$, $a=\lambda =1$ and $b=e$, Theorem \ref{Teo19} yields the
corresponding result which is proven by Agoh and Dilcher \cite{AgohDilcher}.
\end{remark}
\begin{theorem}
The following relationship holds true
\begin{equation*}
\mathfrak{B}_{n}^{(u-v)}(x+y;\lambda ;a,b,b)=\sum_{j=0}^{n}\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) }{\left(
\begin{array}{c}
n+v \\
\end{array
\right) }\mathcal{S}_{v}^{j+v}(x;a,b;\lambda )\mathfrak{B
_{n-j}^{(u)}(y;\lambda ;a,b,b).
\end{equation*}
\end{theorem}
\begin{proof}
We se
\begin{equation}
\left( \frac{\lambda b^{t}-a^{t}}{t}\right) ^{v}b^{xt}\left( \frac{t}
\lambda b^{t}-a^{t}}\right) ^{u}b^{yt}=\left( \frac{t}{\lambda b^{t}-a^{t}
\right) ^{u-v}b^{\left( x+y\right) t}. \label{wc3}
\end{equation
Combining (\ref{w3}) and (\ref{1S}) with the above equation, by using same
calculations with the proof of Theorem \ref{Teo19}, we arrive at the desired
result.
\end{proof}
\section{Application of the Laplace transform to the generating functions
for the generalized Bernoulli polynomials and the generalized array type
polynomials}
In this section, we give an application of the Laplace transform to the
generating function for the generalized Bernoulli polynomials and the
generalized array type polynomials. We obtain interesting series
representation for the families of these polynomials.
By using (\ref{wc3}), we obtai
\begin{eqnarray*}
&&\dsum\limits_{n=0}^{\infty }\mathfrak{B}_{n}^{(u-v)}(\lambda ;a,b,b)\frac
t^{n}}{n!}e^{-t(y-x)\ln b} \\
&=&\dsum\limits_{n=0}^{\infty }\left( \sum_{j=0}^{n}\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) }{\left(
\begin{array}{c}
n+v \\
\end{array
\right) }\mathcal{S}_{v}^{j+v}(x;a,b;\lambda )\mathfrak{B
_{n-j}^{(u)}(\lambda ;a,b,b)\right) \frac{t^{n}}{n!}e^{-ty\ln b}.
\end{eqnarray*
Integrate this equation (by parts) with respect to $t$ from $0$ to $\infty
, we ge
\begin{eqnarray*}
&&\dsum\limits_{n=0}^{\infty }\frac{\mathfrak{B}_{n}^{(u-v)}(\lambda ;a,b,b
}{n!}\dint\limits_{0}^{\infty }t^{n}e^{-t(y-x)\ln b}dt \\
&=&\dsum\limits_{n=0}^{\infty }\left( \frac{1}{n!}\sum_{j=0}^{n}\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) }{\left(
\begin{array}{c}
n+v \\
\end{array
\right) }\mathcal{S}_{v}^{j+v}(x;a,b;\lambda )\mathfrak{B
_{n-j}^{(u)}(\lambda ;a,b,b)\right) \dint\limits_{0}^{\infty }t^{n}e^{-ty\ln
b}dt.
\end{eqnarray*
By using Laplace transform in the above equation, we arrive at the following
Theorem:
\begin{theorem}
\label{TeoL}The following relationship holds true
\begin{equation*}
\dsum\limits_{n=0}^{\infty }\frac{\mathfrak{B}_{n}^{(u-v)}(\lambda ;a,b,b)}
(\ln b^{y-x})^{n+1}}=\dsum\limits_{n=0}^{\infty }\sum_{j=0}^{n}\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) }{\left(
\begin{array}{c}
n+v \\
\end{array
\right) }\frac{\mathcal{S}_{v}^{j+v}(x;a,b;\lambda )\mathfrak{B
_{n-j}^{(u)}(\lambda ;a,b,b)}{\left( \ln b^{y}\right) ^{n+1}}.
\end{equation*}
\end{theorem}
\begin{remark}
When $a=\lambda =1$ and $b=e$, Theorem \ref{TeoL}\ is reduced to the
following result
\begin{equation*}
\dsum\limits_{n=0}^{\infty }\frac{B_{n}^{(u-v)}}{(y-x)^{n+1}
=\dsum\limits_{n=0}^{\infty }\sum_{j=0}^{n}\frac{\left(
\begin{array}{c}
n \\
\end{array
\right) }{\left(
\begin{array}{c}
n+v \\
\end{array
\right) }\frac{S_{v}^{j+v}(x)B_{n-j}^{(u)}}{y^{n+1}}.
\end{equation*}
\end{remark}
\section{Applications the $p$-adic integral to the family of the normalized
polynomials and the generalized $\protect\lambda $-Stirling type numbers}
By using the $p$-adic integrals on
\mathbb{Z}
_{p}$, we derive some new identities related to the Bernoulli numbers, the
Euler numbers, the generalized Eulerian type numbers and the generalized
\lambda $-Stirling type numbers.
In order to prove the main results in this section, we recall each of the
following known results related to the $p$-adic integral.
Let $p$ be a fixed prime. It is known tha
\begin{equation*}
\mu _{q}(x+p^{N}\mathbb{Z}_{p})=\frac{q^{x}}{\left[ p^{N}\right] _{q}}
\end{equation*
is a distribution on $\mathbb{Z}_{p}$ for $q\in \mathbb{C}_{p}$ with $\mid
1-q\mid _{p}<1$, cf. \cite{T. Kim}. Let $UD\left( \mathbb{Z}_{p}\right) $ be
the set of uniformly differentiable functions on $\mathbb{Z}_{p}$. The $p
-adic $q$-integral of the function $f\in UD\left( \mathbb{Z}_{p}\right) $ is
defined by Kim \cite{T. Kim} as follows
\begin{equation*}
\int_{\mathbb{Z}_{p}}f(x)d\mu _{q}(x)=\lim_{N\rightarrow \infty }\frac{1}
[p^{N}]_{q}}\sum_{x=0}^{p^{N}-1}f(x)q^{x},
\end{equation*
wher
\begin{equation*}
\left[ x\right] =\frac{1-q^{x}}{1-q}.
\end{equation*
From this equation, the\textit{\ bosonic} $p$-adic integral ($p$-adic
Volkenborn integral) was considered from a physical point of view to the
bosonic limit $q\rightarrow 1$, as follows (\cite{T. Kim})
\begin{equation}
\int\limits_{\mathbb{Z}_{p}}f\left( x\right) d\mu _{1}\left( x\right)
\underset{N\rightarrow \infty }{\lim }\frac{1}{p^{N}}\sum_{x=0}^{p^{N}-1}
\left( x\right) , \label{M}
\end{equation
wher
\begin{equation*}
\mu _{1}\left( x+p^{N}\mathbb{Z}_{p}\right) =\frac{1}{p^{N}}.
\end{equation*
The $p$-adic $q$-integral is used in many branch of mathematics,
mathematical physics and other areas cf. (\cite{Amice}, \cite{T. Kim}, \cit
{KimSimskJKM}, \cite{Schikof}, \cite{K. Shiratani}, \cite{SimsekJMAA}, \cit
{SimsekADEA}, \cite{srivas18}, \cite{Volkenborn}).
By using (\ref{M}), we have the Witt's formula for the Bernoulli numbers
B_{n}$ as follows
\begin{equation}
\int\limits_{\mathbb{Z}_{p}}x^{n}d\mu _{1}\left( x\right) =B_{n} \label{M1}
\end{equation
cf. (\cite{Amice}, \cite{T. Kim}, \cite{Kim2006TMIC}, \cit
{KimmskimlcjangJIA}, \cite{Schikof}, \cite{Volkenborn}).
We consider the \textit{fermionic} integral in contrast to the convential
bosonic, which is called the fermionic $p$-adic integral on $\mathbb{Z}_{p}$
cf. \cite{Kim2006TMIC}. That i
\begin{equation}
\int\limits_{\mathbb{Z}_{p}}f\left( x\right) d\mu _{-1}\left( x\right)
\underset{N\rightarrow \infty }{\lim }\sum_{x=0}^{p^{N}-1}\left( -1\right)
^{x}f\left( x\right) \label{Mm}
\end{equation
wher
\begin{equation*}
\mu _{1}\left( x+p^{N}\mathbb{Z}_{p}\right) =\frac{(-1)^{x}}{p^{N}}
\end{equation*
cf. \cite{Kim2006TMIC}. By using (\ref{Mm}), we have the Witt's formula for
the Euler numbers $E_{n}$ as follows
\begin{equation}
\int\limits_{\mathbb{Z}_{p}}x^{n}d\mu _{-1}\left( x\right) =E_{n},
\label{Mm1}
\end{equation
cf. (\cite{Kim2006TMIC}, \cite{KimmskimlcjangJIA}, \cite{SimsekADEA}, \cit
{srivas18}).
The Volkenborn integral in terms of the Mahler coefficients is given by the
following Theorem:
\begin{theorem}
\label{TSHiR}Le
\begin{equation*}
f(x)=\sum_{j=0}^{\infty }a_{j}\left(
\begin{array}{c}
x \\
\end{array
\right) \in UD\left( \mathbb{Z}_{p}\right) .
\end{equation*
The
\begin{equation*}
\int\limits_{\mathbb{Z}_{p}}f(x)d\mu _{1}\left( x\right) =\sum_{j=0}^{\infty
}a_{j}\frac{(-1)^{j}}{j+1}.
\end{equation*}
\end{theorem}
Proof of Theorem \ref{TSHiR} was given by Schikhof \cite{Schikof}.
\begin{theorem}
\label{L1
\begin{equation*}
\int\limits_{\mathbb{Z}_{p}}\left(
\begin{array}{c}
x \\
\end{array
\right) d\mu _{1}\left( x\right) =\frac{(-1)^{j}}{j+1}.
\end{equation*
Proof of Theorem \ref{L1} was given by Schikhof \cite{Schikof}.
\end{theorem}
\begin{theorem}
The following relationship holds true
\begin{equation}
B_{m}=\frac{1}{\ln ^{m}b}\sum_{j=0}^{m}(-1)^{j}\frac{j!}{j+1}\mathcal{S
(m,j;1,b;1). \label{1Ss}
\end{equation}
\end{theorem}
\begin{proof}
If we substitute $a=\lambda =1$ in Theorem \ref{T3}, we hav
\begin{equation*}
\left( \ln b^{x}\right) ^{m}=\sum_{j=0}^{m}\left(
\begin{array}{c}
x \\
\end{array
\right) j!\mathcal{S}(m,j;1,b;1).
\end{equation*
By applying the $p$-adic Volkenborn integral with Theorem \ref{L1}\ to the
both sides of the above equation, we arrive at the desired result.
\end{proof}
\begin{remark}
By substituting $b=1$ into (\ref{1Ss}), we have
\begin{equation*}
B_{m}=\sum_{j=0}^{m}(-1)^{j}\frac{j!}{j+1}S(m,j)
\end{equation*
where $S(m,j)$\ denotes the Stirling numbers of the second kind cf. (\cit
{ChanManna}, \cite{http}, \cite{KimChiARXIV}).
\end{remark}
\begin{theorem}
\label{Teo14} The following relationship holds true
\begin{eqnarray*}
&&\sum_{j=0}^{n}\binom{n}{j}\left( \ln a\right) ^{n-j}\left( \ln c\right)
^{j}B_{j}-u(\ln c)^{n}B_{n} \\
&=&\sum_{j=0}^{n}\binom{n}{j}\left( \ln c\right) ^{j}\left( \lambda \left(
\mathcal{H}(u;a,b,c;\lambda )+\ln b\right) ^{n-j}-u\mathcal{H
_{n-j}(u;a,b,c;\lambda )\right) B_{j}.
\end{eqnarray*}
\end{theorem}
\begin{proof}
By using Theorem \ref{t8}, we hav
\begin{eqnarray}
&&\sum_{j=0}^{n}\binom{n}{j}\left( \ln a\right) ^{n-j}\left( \ln c\right)
^{j}x^{j}-u(\ln c)^{n}x^{n} \label{M1a} \\
&=&\sum_{j=0}^{n}\binom{n}{j}\left( \ln c\right) ^{j}x^{j}\left( \lambda
\left( \mathcal{H}(u;a,b,c;\lambda )+\ln b\right) ^{n-j}-u\mathcal{H
_{n-j}(u;a,b,c;\lambda )\right) . \notag
\end{eqnarray
By applying Volkenborn integral in (\ref{M}) to the both sides of the above
equation, we ge
\begin{eqnarray*}
&&\sum_{j=0}^{n}\binom{n}{j}\left( \ln a\right) ^{n-j}\left( \ln c\right)
^{j}\int\limits_{\mathbb{Z}_{p}}x^{j}d\mu (x)-u(\ln c)^{n}\int\limits_
\mathbb{Z}_{p}}x^{n}d\mu (x) \\
&=&\sum_{j=0}^{n}\binom{n}{j}\left( \ln c\right) ^{j}\left( \lambda \left(
\mathcal{H}(u;a,b,c;\lambda )+\ln b\right) ^{n-j}-u\mathcal{H
_{n-j}(u;a,b,c;\lambda )\right) \int\limits_{\mathbb{Z}_{p}}x^{j}d\mu (x).
\end{eqnarray*
By substituting (\ref{M1}) into the above equation, we easily arrive at the
desired result.
\end{proof}
\begin{remark}
By substituting $b=c=e$ and $a=\lambda =1$ into Theorem \ref{Teo14}, we
arrive at the following nice identity
\begin{equation*}
B_{n}=\frac{1}{1-u}\sum_{j=0}^{n}\binom{n}{j}\left( \left( H(u)+1\right)
^{n-j}-uH_{n-j}(u)\right) B_{j}.
\end{equation*}
\end{remark}
\begin{theorem}
\label{Te15} The following relationship holds true
\begin{eqnarray*}
&&\sum_{j=0}^{n}\binom{n}{j}\left( \ln a\right) ^{n-j}\left( \ln c\right)
^{j}E_{j}-u(\ln c)^{n}E_{n} \\
&=&\sum_{j=0}^{n}\binom{n}{j}\left( \ln c\right) ^{j}\left( \lambda \left(
\mathcal{H}(u;a,b,c;\lambda )+\ln b\right) ^{n-j}-u\mathcal{H
_{n-j}(u;a,b,c;\lambda )\right) E_{j}.
\end{eqnarray*}
\end{theorem}
\begin{proof}
Proof of Theorem \ref{Te15} is same as that of Theorem \ref{Teo14}.
Combining (\ref{Mm}), (\ref{M1a}) and (\ref{Mm1}), we easily arrive at the
desired result.
\end{proof}
\begin{remark}
By substituting $b=c=e$ and $a=\lambda =1$ into Theorem \ref{Te15}, we
arrive at the following nice identity
\begin{equation*}
E_{n}=\frac{1}{1-u}\sum_{j=0}^{n}\binom{n}{j}\left( \left( H(u)+1\right)
^{n-j}-uH_{n-j}(u)\right) E_{j}.
\end{equation*}
\end{remark}
\begin{acknowledgement}
The present investigation was supported by the \textit{Scientific Research
Project Administration of Akdeniz University.}
\end{acknowledgement}
|
\section{Introduction}
The last few decades have seen significant progress in the understanding and classification of harmonic maps from surfaces into compact real Lie groups and symmetric spaces.
An important class of harmonic maps are those of \emph{finite type}, which are obtained as the solutions to a pair of ordinary differential equations on a finite dimensional loop algebra. This is a far simpler process than attempting to solve the Laplace-Beltrami equation directly, and so motivates us to determine circumstances under which harmonic maps are of finite type. Similarly, when the target manifold is a $ k $-symmetric space, $ k >2 $,
it is natural to restrict our attention to those harmonic maps which are cyclic primitive and ask when these maps are of finite type. Many papers (e.g. \cite {Hitchin:90, PS:89, Bobenko:91, FPPS:92, BFPP:93, BPW:95, Burstall:95})
have addressed these questions when the target Lie group or ($ k $)-symmetric space is {\em compact}.
We remove the need for this compactness assumption
and in Theorem~\ref {thm:finite} show that all maps from a genus one surface into a $ k $-symmetric space $ G/T $ possessing a Toda frame are of finite type, where $ G $ is any simple real Lie group preserved by a Coxeter automorphism and $ T $ is the corresponding Cartan subgroup. A natural generalisation of the usual 2-dimensional affine Toda field equations provides the integrability condition for the existence of a Toda frame, and so we make contact with classical integrable systems theory.
To determine the spaces $ G/T $ and the harmonic maps into them to which this theory applies we address the following two questions, each of independent interest:
\begin {enumerate}
\item {\em When does a map from a surface into $ G/T $ possess a Toda frame?} and
\item {\em When is $ G $ preserved by a Coxeter automorphism?}
\end {enumerate}
The first of these is answered in Theorem~\ref {theorem:Toda}, where it is proven that a map from a surface into $ G/T $ locally has a Toda frame precisely when it is cyclic primitive and a certain function is constant. Cyclic primitive maps are in particular harmonic and play an analogous role for $ k $-symmetric spaces as harmonic maps do for symmetric spaces.
This and our finite-type result are the natural extensions of results obtained in \cite {BPW:95} in the case when $ G $ is compact. The second question does not arise in the compact situation, since a Coxeter automorphism for a complex simple Lie algebra $\mathfrak{g} ^\mathbb{ C} $ automatically preserves a compact real form $\mathfrak{g} $.
We characterise when a Coxeter automorphism preserves a real form of a complex simple Lie algebra, which is equivalent to the corresponding real Lie group $ G $ being preserved whenever $ G $ simply connected or adjoint. Given simple roots for $\mathfrak{g}^\mathbb{ C} $ spanning a Cartan subalgebra $\t ^\mathbb{ C} $, let $\sigma $ be
the associated Coxeter automorphism and $\Theta $ a Cartan involution with respect to $\mathfrak{g} $ that preserves $\t =\mathfrak{g}\cap\t ^\mathbb{ C} $.
Then $\sigma $ preserves $\mathfrak{g} $ if and only if $\Theta $ defines a permutation of the extended Dynkin diagram, so in particular whenever $\t $ is a maximally compact Cartan subalgebra (Proposition~\ref{prop:Coxeter}).
In Theorem~\ref{thm:3.2}
we prove that all involutions of the extended Dynkin diagram for a simple complex Lie algebra $\mathfrak{g} ^\mathbb{ C} $ arise from a Cartan involution for some real form $\mathfrak{g} $.
Harmonic maps from surfaces into Lie groups and symmetric spaces arise naturally in many geometric and physical problems. On the geometric side, strong motivation comes from the study of surfaces with particular curvature properties. For example, minimal surfaces are described by conformal harmonic maps and both constant mean curvature and Willmore surfaces are characterised by having harmonic Gauss maps into particular symmetric spaces. From the physics viewpoint, these harmonic maps are interesting because of their relationship with the appropriate Yang-Mills equations and non-linear sigma-models. Indeed the harmonic map equations on a Riemann surface are precisely the reduction of the Yang-Mills equations on $\mathbb{ R} ^ {2, 2} $ obtained by considering solutions invariant under translation in the directions of negative signature. Classical solutions of sigma-models are given by harmonic maps into (non-compact) as pseudo-Riemannian manifolds. In \cite{CT:12} we study an explicit example, namely harmonic tori in de Sitter spaces $ S ^ {m}_1 $. In particular we apply the theory of this paper to the superconformal such maps with globally defined harmonic sequence to see that they may all be obtained by integrating a pair of commuting vector fields on a finite-dimensional vector space. It follows that all Willmore tori in $ S ^ 3 $ without umbilic points may be obtained in this simple way.
The structure of this paper is as follows. In section~\ref {symmetric} we give the general theory for harmonic maps of surfaces into symmetric spaces and for primitive maps into $ k $-symmetric spaces when the relevant Lie group $ G $ is equipped with a bi-invariant pseudo-metric.
The question of when a Coxeter automorphism preserves the real form of the complex simple Lie algebra is addressed in section ~\ref{dynkin} in terms of Cartan involutions and extended Dynkin diagrams.
Section~\ref {Toda} contains the relationship with the affine Toda field equations and the finite type result is proven in section~\ref {finite}.
It is a pleasure to thank Anthony Henderson for helpful conversations regarding the Lie-theoretic results of section~\ref {dynkin}.
\section{Finite type maps into symmetric spaces}\label{symmetric}
The fact that a harmonic map from a surface to a Lie group corresponds to a loop of flat connections \cite {Pohlmeyer:76, Uhlenbeck:89} is the fundamental observation that enables one to apply integrable systems techniques to the study of these maps. The Cartan map $ G/H\rightarrow G $ from a symmetric space to the relevant Lie group is well-known to be a totally geodesic immersion when $ G $ is compact and equipped with a bi-invariant Riemannian metric. The composition of a harmonic map with a totally geodesic one is again harmonic, so this enables harmonic maps into symmetric spaces to be studied using the same tools as those into Lie groups, and in particular in terms of a loop of flat connections. We show in Theorem~\ref{thm2.1}
that when $ G $ has merely a bi-invariant pseudo-metric that the Cartan map is again a totally geodesic immersion. In particular all reductive Lie groups possess a bi-invariant pseudo-metric. We can hence study harmonic maps into $ G/H $ using integrable systems methods regardless of whether $ G $ is compact.
Let $ G $ be a semisimple Lie group. Recall that a homogeneous space $ G/H $ is a {\it $ k $-symmetric space} ($ k >1 $) if there is an automorphism $\tau: G\rightarrow G $ of order $ k $ such that
\[
(G ^\tau)_0\subset H\subset G ^\tau
\] where $ G ^\tau $ denotes the fixed point set of $\tau $, and $ (G ^\tau)_0 $ the identity component of $ G ^\tau $. When $ k = 2 $, we say that $ G/H $ is a {\it symmetric space}.
We have the induced action
\begin {align*}
\tau: G/H &\rightarrow G/H\\
gH &\mapsto\tau (g) H.
\end {align*}
We write $\tau $ also for the induced automorphism of $\mathfrak{g} $ and note the $\mathbb{ Z}_k $-grading
\[
\mathfrak{g} ^\mathbb{ C} =\bigoplus_{j = 0} ^ {k -1}\mathfrak{g} ^\tau_j,\; [\mathfrak{g} ^\tau_j,\mathfrak{g} ^\tau_l ]\subset\mathfrak{g} ^\tau_{j + l},
\]
where $\mathfrak{g} ^\tau_j $ denotes the $ e ^ {j\frac {2\pi i} {k} } $-eigenspace of $\tau $.
We shall be interested in harmonic maps from a Riemann surface $\Sigma $ into a symmetric space $ G/H $. When $ G $ is compact, the Killing form on $\mathfrak{g} $ induces a bi-invariant metric on $ G/H $ and the harmonic map equations for $ f:\Sigma\rightarrow G/H $ may either be calculated directly \cite{Wood:94}, using Noether's Theorem \cite {Rawnsley:84},
or by composing $ f $ with the Cartan map $ G/H\rightarrow G $, which is well-known in this case to be a totally geodesic immersion \cite {CE:75}. Recall here that the Cartan map of a symmetric space is given by
\begin {align*}
\iota:\;\; & G/H\rightarrow G\\
& gH\mapsto \tau (g ) g^{-1}.
\end {align*}
We suppose merely that $ G $ has a bi-invariant pseudo-metric. Then analogous computations hold; in particular we can reduce the problem to studying harmonic maps into the Lie group $ G $ due to the following result.
\begin {theorem}\label{thm2.1}
Let $ G $ be a semisimple Lie group with bi-invariant pseudo-metric $\langle \cdot,\cdot \rangle $ and $ G/H $ a symmetric space with respect to the involution $\tau: G\rightarrow G $.
Then $\iota: gH\mapsto \tau (g)g ^ {- 1} $ is a totally geodesic immersion $ G/H\rightarrow G $ .
If $ H = G ^\tau $, then $\iota$ is additionally an embedding.
\end {theorem}
\noindent Let us call a Lie group $ G $ {\em reductive} if its Lie algebra $\mathfrak g $ is reductive, that is has radical equal to its centre. Then $\mathfrak g $ may be written as the direct sum of a semisimple Lie algebra and an abelian one. On the semisimple Lie algebra the Cartan-Killing form is non-degenerate, whilst on the abelian algebra any bilinear form is invariant under the adjoint action of the group. Combining these we obtain the existence of a bi-invariant pseudometric on any reductive Lie group, and hence the above theorem in particular applies when $ G $ is reductive.
\begin {proof}{\em $\iota$ is an immersion:} Suppose $d\iota_{gH} (\gamma' (0)) = 0 $ for some smooth path $\gamma $ in $ G/H $ with $\gamma (0) = gH $. Take a lift $\tilde\gamma $ of $\gamma $ to $ G $ with $\tilde\gamma (0) = g $ and write $\pi: G\rightarrow G/H $ for the projection. Then
\[
0 =
\left.\dfrac{d}{dt}\right|_{t=0}\left (\tau\left(\tilde\gamma(t)\right)\left(\tilde\gamma(t)\right)^{-1}\right) = d\tau_g (\tilde\gamma' (0)) g^ {- 1} -\tau (g) g^ {- 1}\tilde\gamma' (0) g^ {- 1},
\]
so
\[
d\tau_e (g^ {- 1}\tilde\gamma' (0)) =\tau (g^ {- 1}) d\tau_g (\tilde\gamma' (0)) = g^ {- 1}\tilde\gamma' (0)
\]
and $\gamma'(0)$ is zero in $T_{g H}(G/H)$ so $d\iota_{g H}$ is injective.
{\em $\iota$ is totally geodesic:}
Let $\nabla ^l $ denote the connection on $ G $ obtained by trivialising $ TG $ by left translation, and similarly $\nabla ^ r $ that induced from trivialising by right translation. A computation shows that
$\nabla ^ r =\nabla ^ l +\mathrm{ad}_{g^ {- 1} dg} $ and hence
\[
\nabla =\frac 12 (\nabla ^ l +\nabla ^ r)
\]
is the Levi-Civita connection of the pseudo-metric $\langle \cdot ,\cdot \rangle $.
Denote by $\exp:\mathfrak g\rightarrow G $ the Lie-theoretic exponential map, and by $e $
the differential-geometric exponential map associated to the Levi-Civita connection $\nabla $.
Note that as in the definite case, for each $ X\in\mathfrak g $
the map
\begin {align*}
\gamma_X: \mathfrak{g}&\rightarrow G\\
t&\mapsto e ^ {tX}
\end {align*}
\noindent is a geodesic, i.e.
$\nabla_{\gamma'_X}\gamma'_X = 0 $, so $ \exp $ and $e $ agree on the domain of $e $. Since the pseudo-metric is bi-invariant, we conclude that the geodesics through $ g\in G $ are locally of the form $\gamma (t) = g e ^ {tX} $.
Denote by $\mathfrak{m} $ the $(-1)$-eigenspace of $\tau:\mathfrak{g}\rightarrow\mathfrak{g} $, and note that $\mathfrak{g} =\mathfrak{h}\oplus \mathfrak{m} $, where $\mathfrak{h} $ is
the Lie algebra of $ H $. The lift $\tilde\gamma (t) = g e ^ {tX} H $ is horizontal, in the sense that $\tilde\gamma' (t)\in g e ^ {tX}\mathfrak{m} $. Thus the geodesics in $ G/H$ through $ g H $ are locally of the form $\tilde\gamma (t) = g e ^ {tX} H $. Since
\[
\iota(g e ^ {tX} H) = g e ^ {tX}\tau (e ^ {- tX})\tau (g^ {- 1}) = g e ^ {2 tX}\tau (g^ {- 1}) = g\tau (g^ {- 1}) e ^ {t\tau (g) X\tau (g^ {- 1})}
\]
is again a geodesic, we conclude that $\iota$ is totally geodesic.
{\em If $ H = G ^\tau $, then $\iota$ is an embedding:} In this case if $\iota(g_1H) =\iota(g_2H) $, then $ g_1 ^ {- 1} g_2 =\tau (g_1 ^ {- 1} g_2) $, and so $ g_1 ^ {- 1} g_2\in H $, and thus $\iota$ is injective.
\end {proof}
Let $ F:U\rightarrow G $ be a smooth lift of $ f:U\rightarrow G/H $ on some simply connected $ U\subset\Sigma $, where we assume henceforth that $ G $ is semisimple and has a bi-invariant pseudo-metric (we will later restrict our attention to simple such $ G $.). By the above theorem, $ f $ is harmonic if and only if $\iota\circ f $ is. The Maurer-Cartan form on $ G $ is the unique left-invariant $\mathfrak{g}$-valued 1-form which acts as the identity on $\mathfrak{g} $. We denote it by $\omega $, and note that if $ G $ is a linear group, then $\omega= g ^ {- 1} dg $. We will use this notation throughout even in the non-linear case.
Write $ \tilde {f} = \iota\circ f $ and $\Phi = \tilde {f} ^*(\omega) = \tilde {f} ^ {- 1} d\tilde {f} $. For any smooth $ \tilde {f} $, the form $\Phi $ satisfies the zero-curvature condition
\begin {equation}
d\Phi +\frac 12 [\Phi\wedge\Phi ] = 0,\label {eq:MC}
\end {equation}
known as the Maurer-Cartan equation. Recall that for vector fields $ X, Y $,
\[
[\Phi\wedge\Phi] (X, Y) = 2 [\Phi,\Phi] (X, Y) = [\Phi (X),\Phi (Y)].
\]
The condition that the map $ \tilde {f}:\Sigma\rightarrow G $ is harmonic can be written as
\begin {equation}
d*\Phi = 0.\label {eq:harmonic}
\end {equation}
Noting that $ \tilde {f} =\tau (F) F^ {- 1} $, we have
\begin {equation}
\label {eq:composesymmetric}
\Phi = F\left (\tau (F)^ {- 1} d (\tau (F)) - F^{-1}d F \right) F^ {- 1} = -2\mathrm{Ad}_F (\varphi_{\mathfrak{m}}),
\end {equation}
where $\varphi =\varphi _{\mathfrak{h}} +\varphi _{\mathfrak{m}}$ is the decomposition of $\varphi : = F^ {- 1} d F $ into the eigenspaces of $\tau $. Then \eqref {eq:harmonic} becomes
\begin {equation}\label {eq:harmonic1st}
0 = d (\mathrm{Ad}_F (*\varphi _{\mathfrak{m}})) =\mathrm{Ad}_F (d*\varphi _{\mathfrak{m}} + [\varphi \wedge*\varphi_{\mathfrak{m}} ])
\end {equation}
or equivalently,
\begin {equation}
d*\varphi_{\mathfrak{m}} + [\varphi\wedge*\varphi_{\mathfrak{m}} ] = 0.\label{eq:harmonicsymmetric}
\end{equation}
One can also compute the harmonic map equations directly for $ f $. Writing $ [\mathfrak{m}] $ for the subbundle of $ G/H\times\mathfrak{g} $ whose fibre at $ g\cdot x $ is $\mathrm{Ad}_g (\mathfrak{m}) $, we have an isomorphism $ [\mathfrak{m}]\cong {T (G/H)}] $ given by
\begin {align*}
[\mathfrak{m} ]_y &\rightarrow T_y G/H\\
Y &\mapsto\left.\frac {d} {dt}\right|_{t = 0} e ^ {tY}\cdot y.
\end {align*}
The inverse of this isomorphism defines a $\mathfrak{g} $-valued 1-form $\theta $ on the symmetric space $ G/H $, which we term its Maurer-Cartan form. Then \cite {Rawnsley:84} $ f $ is harmonic if and only if
\[
d*(f ^*\theta) = 0
\]
and using that
\[
f ^*\theta =\mathrm{Ad}_F (\varphi_{\mathfrak{m}})
\]
we recover \eqref {eq:harmonic1st}.
Write $\varphi'_\mathfrak{m} +\varphi_\mathfrak{m}'' $ for the decomposition of $\varphi_\mathfrak{m} $ into $ dz $ and $ d\bar z $ parts. Since $ [\mathfrak{m},\mathfrak{m} ]\subset\mathfrak{h} $, a straightforward computation shows \eqref {eq:MC} and \eqref {eq:harmonicsymmetric} are equivalent to the requirement that for each $\lambda\in S ^ 1 $, the form
\begin {equation}\label {eq:form}
\varphi _\lambda =\lambda\varphi'_\mathfrak{m} +\varphi_\mathfrak{h} +\lambda^ {- 1}\varphi''_\mathfrak{m}
\end {equation}
satisfies the Maurer-Cartan equation
\begin {equation}
d\varphi_\lambda +\frac 12 [\varphi_\lambda\wedge\varphi_\lambda ] = 0.\label {eq:flat}
\end {equation}
Some solutions to \eqref {eq:flat} can be obtained simply by solving a pair of commuting ordinary differential equations on a finite-dimensional loop algebra. These unusually simple solutions are said to be of finite type.
Let $ G/K $ be a $ k $-symmetric space for $ k >2 $ and $\tau $ the corresponding $ k $th order involution. As we shall now explain when mapping into a $ k $-symmetric space for $ k >2 $ it is natural to restrict our attention to a subclass of harmonic maps consisting of those which are primitive, a notion that we now define. Again we have the reductive splitting
\[
\mathfrak{g} =\mathfrak{k}\oplus\mathfrak{p}
\]
with
\[
\mathfrak{p} ^\mathbb{ C} =\bigoplus_{j = 1} ^ {k -1}\mathfrak{g}_j ^\tau,\qquad \mathfrak {k} ^\mathbb{ C} =\mathfrak{g}_0 ^\tau.
\]
Similarly to before we may define the Maurer-Cartan form $\theta $ of the $ k $-symmetric space $ G/K $ when $ k >2 $. For any smooth lift $ F: U\rightarrow G $ of $ \psi: U\rightarrow G/K $, writing $\varphi = F ^*\omega $ we have
\[
\psi ^*\theta =\mathrm{Ad}_F\varphi_{\mathfrak{p}}.
\]
We say that a smooth map $ \psi$ of a surface $\Sigma $ into $ G/K $ is {\em primitive} if the image of $ \psi ^*\theta' $ is contained in $ [ \mathfrak{g}_1 ] $. Equivalently, it is primitive precisely when
$\varphi ' = F ^ {- 1}\partial F $ takes values in $\mathfrak{g}_0 ^\tau\oplus\mathfrak{g} _1^\tau $. Using that $ [\mathfrak{g}_1^\tau,\mathfrak{g}_{-1} ^\tau]\subset\mathfrak{g}_0^\tau $, the Maurer-Cartan equation for $\varphi $ decomposes into $\mathfrak{g}_1^\tau $, $\mathfrak{g}_0^\tau $ and $\mathfrak{g}_{-1} ^\tau $ components as
\begin {align}
d\varphi'_\mathfrak{p} + [\varphi _{\mathfrak {k}}\wedge\varphi '_\mathfrak{p}] & = 0\label {eq:MCg1}\\
d\varphi_{\mathfrak {k}} + \frac 12 [\varphi _{\mathfrak {k}}\wedge\varphi_{\mathfrak {k}}] + [\varphi '_\mathfrak{p}\wedge\varphi''_\mathfrak{p}] & = 0\nonumber\\
d\varphi''_\mathfrak{p} + [\varphi _{\mathfrak {k}}\wedge\varphi ''_\mathfrak{p}] & = 0.\nonumber
\end {align}
From these equations one easily verifies that primitive maps are in particular harmonic. Moreover \cite{BP:94} if $ G/H $ is a symmetric space with $ K\subset H $ and the corresponding reductive splitting preserved under $\tau $, then the projection of $ \psi:\Sigma\rightarrow G/K $ into $ G/H $ is harmonic. An analogous calculation to that above shows that on simply connected subsets $ U\subset\Sigma $, a primitive map $ \psi: U\rightarrow G/K $ is equivalent to a loop
\begin {equation}
\label {eq:primitiveflat}
\varphi_\lambda =\lambda \varphi '_\mathfrak{p} +\varphi_{\mathfrak{k}} +\lambda^ {- 1}\varphi''_\mathfrak{p},\quad\lambda\in S ^ 1
\end {equation}
of $\mathfrak{g} $-valued 1-forms each satisfying the Maurer-Cartan equation. We see then that both harmonic maps into symmetric spaces and primitive maps into $ k $-symmetric spaces are governed by the same equation \eqref{eq:flat} so we turn now to the question of constructing solutions to this equation.
Let $\Omega G $ be the loop group
$
\Omega G =\{\gamma: S ^ 1\rightarrow G \}
$ with corresponding loop algebra
$
\Omega\mathfrak{g}: =\{\xi: S ^ 1\rightarrow\mathfrak{g}\} $ , where the loops are assumed real analytic without further comment.
We use $\Omega\mathfrak{g} ^\mathbb{ C} $ to denote loops in the complexified Lie algebra $\mathfrak{g} ^\mathbb{ C} $.
For studying maps into $ k $-symmetric spaces it is helpful to consider the twisted loop group
\[
\Omega ^\tau G =\{\gamma: S ^ 1\rightarrow G:\gamma (e ^{\frac {2\pi i} {k}\lambda}) =\tau (\gamma (\lambda))\}
\]
and corresponding twisted loop algebra $\Omega ^\tau\mathfrak{g} $ along with its complexification $\Omega ^\tau\mathfrak{g} ^\mathbb{ C} $.
The (possibly doubly infinite) Laurent expansion
\[
\xi(\lambda) =\sum_{j} \xi_j \lambda ^ j,\quad\xi_j\in\mathfrak{g} ^\tau_j\subset\mathfrak{g} ^\mathbb{ C},\quad\Phi_{- j} =\bar\Phi_j
\]
allows us to filtrate $\Omega ^\tau\mathfrak{g} ^\mathbb{ C} $ by finite-dimensional subspaces
\[
\Omega ^\tau_d =\{\xi\in\Omega\mathfrak{g}\mid \xi_j = 0\text { whenever }\left|j\right| >d\}.
\]
Fix a Cartan subalgebra $\mathfrak t$ of $\mathfrak{g} $ such that $\mathfrak t\subset\mathfrak k $
and recall that a non-zero $\alpha \in (\mathfrak{t}^\mathbb{ C})^*$ is a {\em root} with corresponding {\em root space} $\mathcal{G}^{\alpha}\subset\mathfrak{g} ^\mathbb{ C} $
if $[X_1, X_2 ]=\alpha(X_1) X_2 $ for all $X_1\in \mathfrak{t}$ and $ X_2\in \mathcal{G}^{\alpha}$.
We denote the set of roots by $\Delta$ and employ the same notation for the root system formed by considering $\Delta $ as a subset of $ (\mathfrak {t} ^\mathbb{ C}) ^*$.
Choose a set of {\em simple roots}, that is a subset $\{ \alpha_1, \ldots, \alpha_N \}$ of $\Delta$ such that every root $\alpha\in \Delta$ can be written uniquely as
$$\alpha=\sum_{j=1}^{N} m_j \alpha_j,$$ where the $m_j$ are either all positive integers or all negative integers.
The {\em height } of $\alpha$ is $h(\alpha) =\sum_{j = 1} ^ N m_j$ and the root(s) of maximal height are called {\em highest root(s)} whilst those of minimal height are termed {\em lowest root(s)}.
We similarly define the root spaces
of $\mathfrak k ^\mathbb{ C} $. Let $\mathfrak n $ be the nilpotent algebra consisting of the positive root spaces of $\mathfrak k ^\mathbb{ C} $ with respect to a choice of simple roots and consider the resulting Iwasawa decomposition
\begin{equation}\label {eq:Iwasawa}
\mathfrak{k}^\mathbb{ C} =\mathfrak{n}\oplus\mathfrak{t}^\mathbb{ C}\oplus\bar{\mathfrak {n}}
\end {equation}
of $\mathfrak {k} ^\mathbb{ C} $.
Then for $\eta\in\mathfrak k ^\mathbb{ C} $ and a local coordinate $ z $ on $\Sigma $, decomposing according to \eqref{eq:Iwasawa} we have
\[
(\eta dz)_{\mathfrak h} = r (\eta) dz +\overline {r (\eta)} d\bar z
\]
where $r:\mathfrak{k}^\mathbb{ C}\rightarrow\mathfrak{k}^\mathbb{ C} $ is defined by
\begin {equation}\label {eq:r}
r (\eta) =\eta_{\bar{\mathfrak{n}}} +\frac 12\eta_{\mathfrak k}.
\end {equation}
The key observation here is that for simply-connected coordinate neighbourhood $ U\subset\Sigma $, if $\xi: U\rightarrow\Omega ^\tau_d $ satisfies
\begin {equation}\label{eq:flows}
\dfrac {\partial\xi} {\partial z} = [\xi,\lambda\xi_d + r (\xi_{d -1})]
\end{equation}
then
\[
\varphi_\lambda = (\lambda\xi_d + r (\xi_{d -1})) d z + (\lambda^ {- 1}\xi_{- d} +\overline {r (\xi_{d -1})}) d\bar z
\]
satisfies the Maurer-Cartan equation \eqref{eq:flat} (c.f. \cite{BP:94}, Theorem 2.5). The equation
\[
\frac 12 ( X (\xi) - iY (\xi)) = (\lambda\xi_d + r (\xi_{d -1}))
\]
defines vector fields $ X, Y $ on $\Omega_d $. A straightforward computation shows that these vector fields commute and so finding solutions to \eqref {eq:flows} is merely a matter of solving a pair of commuting ordinary differential equations. This yields a rather special class of solutions to the Maurer-Cartan equations \eqref {eq:flat} and hence of harmonic maps to symmetric spaces and primitive maps to $ k $-symmetric spaces, $ k >2 $.
The flows of $ X, Y $ are easily seen to evolve on spheres in $\Omega_d $. When $ G $ is compact, so are these spheres and hence $ X, Y $ are complete and for any initial condition the differential equation \eqref {eq:flows} has a unique solution on $ U $. However when $ G $ is non-compact the completeness of $ X, Y $ is not guaranteed.
\begin {definition}
A harmonic map $ f:\Sigma\rightarrow G/H $ to a symmetric space or a primitive map $ \psi:\Sigma\rightarrow G/K $ to a $ k $-symmetric space, $ k >2 $ is said to be of {\em finite type} if it has a lift $ F:\Sigma\rightarrow G $ for which there exists
a smooth map $\xi: \mathbb{ R} \to\Omega_d ^\tau\mathfrak{g}$ satisfying
\begin {equation}\label {eq:lax}
d\xi = [\xi,\varphi _\lambda ]
\end {equation}
and
\begin {equation}\label {eq:adapted}
\varphi_\lambda = (\lambda\xi_d + r (\xi_{d -1})) d z + (\lambda^ {- 1}\xi_{- d} +\overline {r (\xi_{d -1})}) d\bar z.
\end {equation}
Here $\varphi_\lambda $ and $ r $ are defined in \eqref {eq:primitiveflat}
and \eqref {eq:r} for the primitive case and in \eqref{eq:form} and the obvious analogue to \eqref {eq:r} for the harmonic case.
\end {definition}
We introduce some terminology for later use. A {\em formal Killing field} for $ f $ or $ \psi $ is a smooth map $\xi:\Sigma\rightarrow\Omega ^\tau\mathfrak g $ satisfying the Lax equation \eqref{eq:lax}. When $\xi $ takes values in some $\Omega_d$ it is termed a {\em polynomial Killing field} of degree $ d $ and when it additionally satisfies \eqref{eq:adapted} it is an {\em adapted polynomial Killing field}.
When the automorphism $\tau: \mathfrak{g} ^\mathbb{ C}\rightarrow \mathfrak{g} ^\mathbb{ C} $ is of the form $\tau =\mathrm{Ad}_{\exp M}$ for some
$ M\in\mathfrak t ^\mathbb{ C} $ where $\mathfrak t $ is a Cartan subalgebra of $\mathfrak g $,
then we can express the eigenspaces $\mathfrak{g} ^\tau_j $ of $\tau $ in terms of root spaces.
Given our chosen set of simple roots $\alpha _j $, denote by $ \eta_j$ the corresponding dual basis of $\mathfrak{t} ^\mathbb{ C}$.
For any root $\alpha=m_1\alpha_1 + \ldots m_N \alpha_N$, smooth map $s_j:\Sigma \to \mathbb{ C}$ and root vector $R_\alpha\in\mathcal{G} ^\alpha $, a straightforward computation shows that
\begin {equation}
\operatorname{Ad}_{ \exp(s_1 \eta_1+\ldots s_N \eta_N)} R_\alpha = \exp(m_1 s_1+\ldots m_Ns_N)R_\alpha.\label {eq:roots}
\end {equation}
Note that $ \exp(m_1 s_1+\ldots m_Ns_N)$ is a scalar function.
Given $\tau = \operatorname{Ad}_{ \exp (\frac{2\pi i}{k}(\sum s_j \eta_j))} $ we have
$$\mathfrak{g}^\tau_l = \Span \{R_\alpha| \alpha=\sum_{j = 1} ^ N m_j\alpha_j, \sum_{j = 1} ^ N s_jm_j=l\bmod(k)\}.$$
In particular if we let $k-1$ denote the maximal height of a root of $\mathfrak{g} ^\mathbb{ C} $ and
suppose
\begin {equation}\label {eq:Coxeterdefinition}\sigma:=\operatorname{Ad}_{ \exp ( \frac {2\pi i } {k}\sum_{j = 1} ^ N \eta_j)}\,,
\end {equation}
then $\sigma$ is of order $k$ and from \eqref {eq:roots} it acts on the root spaces by
\begin {equation}\label {eq:Coxeter}
\sigma(R_\alpha) = \exp\left(\frac{2\pi i h(\alpha)}{k}\right)R_\alpha.
\end {equation}
We recognise the inner automorphism $\sigma$ as the Coxeter automorphism associated to the identity transformation of the simple roots \cite {BD:81}. It plays an important role here because when it preserves the real Lie group $ G $,
it allows us to view $G/T$ as a $k$-symmetric space for which $\mathfrak{g} ^\sigma_1 $
is the sum of the simple and lowest root spaces. Here $ T $ is a Cartan subgroup with Lie algebra $\mathfrak t $. Furthermore since $ K = T $ in this case, the map $ r $ described in \eqref {eq:r} is simply multiplication by $\tfrac {1} {2} $ and so the adapted polynomial Killing field condition \eqref {eq:adapted} simplifies. Taking this $ N $-symmetric space structure on $ G/T $, we say that a smooth map $ \psi:\Sigma\rightarrow G/T $ is \emph{cyclic primitive} if it is primitive and satisfies the condition that the image of $ \psi ^*\theta' $ contains a cyclic element. Writing $\alpha_0 $ for the lowest root, an element in $\left (\bigoplus_{j = 0} ^ N \mathcal G ^{\alpha_j}\right) $ is \emph{cyclic} if its projection to each of the root spaces $ \mathcal G ^{\alpha_0}, \mathcal G ^{\alpha_1},\ldots , \mathcal G ^{\alpha _N} $ is non-zero. We henceforth assume that $ G $ is simple in order to guarantee the uniqueness of the lowest root (that is, we assume that $ G $ is connected and $\mathfrak{g} $ is simple).
\section {Extended Dynkin diagrams and Cartan involutions}\label{dynkin}
We now ascertain the $ k $-symmetric spaces to which our theory will apply. That is, we give conditions under which a choice of real form $\mathfrak{g} $ of a simple complex Lie algebra $\mathfrak{g} ^\mathbb{ C} $, Cartan subalgebra $\mathfrak t ^\mathbb{ C} $ and simple roots $\alpha_j $
yield a Coxeter automorphism $\sigma $ which preserves the real Lie algebra $\mathfrak{g} $. When $ G ^\mathbb{ C} $ is a simply connected or adjoint simple Lie group with Lie algebra $\mathfrak{g} ^\mathbb{ C} $, this ensures that the Coxeter automorphism preserves the real group $ G $.
Let $\;\bar{}\;$
denote the complex conjugation of $\mathfrak{g} ^\mathbb{ C} $ corresponding to the real form $\mathfrak{g} $. Define the conjugate of a root $\alpha $ by
\[
\bar\alpha (X) =\overline {\alpha (\bar X)}.
\]
Then from \eqref {eq:Coxeter} we see that the condition for the Coxeter automorphism $\sigma $ to preserve $\mathfrak{g} $ is that for all roots $\alpha $, the height $ h (\alpha $) satisfies
\[
h (\bar\alpha) = - h (\alpha)\bmod k,
\]
or equivalently that for
$ j = 1,\ldots, N $ we have
\begin {equation*}\label {eq:simplereality}
\bar\alpha_j\in\{-\alpha_0,\ldots, -\alpha_N\}.
\end {equation*}
We will now use a Cartan involution to express this reality condition in terms of the extended Dynkin diagram for $\alpha_0,\ldots,\alpha_N $. A Cartan involution is an involution $\Theta $ of $\mathfrak{g} $ such that
\[
\langle X, Y \rangle _{\Theta} = - \langle X,\Theta (Y) \rangle
\]
is positive definite, where $ \langle \cdot ,\cdot \rangle $ denotes the Killing form. Using complex-linearity, $\Theta $ extends to an involution of $\mathfrak{g} ^\mathbb{ C} $.
We may \cite[Prop.\! 6.59]{Knapp:02}
choose a Cartan involution $\Theta $ which preserves the Cartan subalgebra $\mathfrak t $.
\begin {proposition}\label {prop:Coxeter}
Let $\mathfrak{g} $ be a real simple Lie algebra, $\t $ a Cartan subalgebra and $\Theta $ be a Cartan involution preserving $\mathfrak t $. Choose simple roots $\alpha_1,\ldots,\alpha_N $ for the root system $\Delta (\mathfrak{g} ^\mathbb{ C},\t ^\mathbb{ C}) $ and let $\sigma $ be the corresponding Coxeter automorphism of $\mathfrak{g} ^\mathbb{ C} $ defined in \eqref {eq:Coxeterdefinition}. Then the following are equivalent:
\begin {enumerate}
\item $\sigma $ preserves the real form $\mathfrak{g} $,
\item $\sigma $ commutes with $\Theta $,
\item $\Theta $ defines a permutation of the extended Dynkin diagram for $\mathfrak{g} ^\mathbb{ C} $ consisting of the usual Dynkin diagram augmented with the lowest root $\alpha_0 $.
\end {enumerate}
\end {proposition}
\begin {proof} Write $\mathfrak t =\mathfrak l\oplus \mathfrak p $, where $\mathfrak l $, $\mathfrak p $ are respectively the $ (+ 1 ) $-eigenspace and $ (- 1) $-eigenspace of the action of $\Theta $ on $\mathfrak t $.
Then \cite[Cor. 6.49] {Knapp:02} all roots $\alpha $ are real on $\mathfrak p $ and imaginary on $\mathfrak l $, and defining the action of $\Theta $ on roots by $\Theta (\alpha) (X) =\alpha (\Theta (X)) $ we have that
\[
\Theta (\alpha) = -\bar\alpha\quad\text { for all roots }\alpha.
\]
If $ R_\alpha $ is a root vector for $\alpha $, then $\bar R_\alpha $ is a root vector for $\bar\alpha $ and $\Theta (R_\alpha) $ is a root vector for $\Theta (\alpha) $. We assume that our root vectors are chosen so that
\[
R_{\bar\alpha} =\bar R_\alpha
\]
and write $ R_{\Theta (\alpha)} = c_\alpha\Theta (R_\alpha) $. Then using \eqref {eq:Coxeter}, a straightforward computation shows that $\sigma\circ\Theta (R_\alpha) =\Theta\circ\sigma (R_\alpha) $ if and only if $\sigma (\bar R_{-\alpha}) =\overline {\sigma (R_{-\alpha})} $,
proving the equivalence of conditions (1) and (2) above.
The Cartan involution $\Theta $ commutes with $\sigma $ if and only if for all roots $\alpha $, the height function $ h $ satisfies
\[
h (\Theta (\alpha))\equiv h (\alpha)\bmod k,
\]
or equivalently when $\Theta $ defines a permutation of
$\alpha_0,\alpha_1,\ldots,\alpha_N $. All automorphisms of a Lie algebra preserve the Killing form and hence a Cartan involution $\Theta $ as above defines a permutation of the extended Dynkin diagram
and we see the equivalence of conditions (2) and (3).
\end {proof}
We next show that every involution of the extended Dynkin diagram for $\Delta (\mathfrak{g} ^\mathbb{ C},\t ^\mathbb{ C}) $ does indeed arise from a Cartan involution for some real form $\mathfrak{g} $ with $\Theta $-stable Cartan subalgebra $\t =\mathfrak{g}\cap\t ^\mathbb{ C} $.
A $\Theta $-stable Cartan subalgebra $\t $ of $\mathfrak{g} $ is maximally compact if and only if $\Theta $ preserves a set of simple roots for the root system $\Delta (\mathfrak{g},\t)$ \cite [p 387] {Knapp:02}. Hence when $\t $ is maximally compact, a Coxeter automorphism $\sigma $ must stabilise the real form $\mathfrak{g} $. (In particular, all Cartan subalgebras of a compact real form $\mathfrak{g} $ are maximally compact.) The more interesting case then is when the Cartan subalgebra $\t $ is not maximally compact, which corresponds to the involution of the extended Dynkin diagram acting nontrivially on the lowest root $\alpha_0 $.
\begin{theorem}\label{thm:3.2} Every involution of the extended Dynkin diagram for a simple complex Lie algebra $\mathfrak{g} ^\mathbb{ C} $ is induced by a Cartan involution of a real form of $\mathfrak{g} ^\mathbb{ C} $.
More precisely, let $\mathfrak{g}^\mathbb{ C} $ be a simple complex Lie algebra with Cartan subalgebra $\t ^\mathbb{ C} $ and choose simple roots $\alpha_1,\ldots,\alpha_N $ for the root system $\Delta (\mathfrak{g} ^\mathbb{ C},\t ^\mathbb{ C}) $. Given an involution $\pi $ of the extended Dynkin diagram for $\Delta $,
there exists a real form $\mathfrak{g} $ of $\mathfrak{g} ^\mathbb{ C} $ and a Cartan involution $\Theta $ of $\mathfrak{g} $ preserving $\t =\mathfrak{g}\cap\t ^\mathbb{ C} $ such that $\Theta $ induces $\pi $ and $\t $ is a real form of $\t ^\mathbb{ C} $. The Coxeter automorphism $\sigma $ determined by $\alpha_1,\ldots,\alpha_N $ preserves the real form $\mathfrak{g} $.
\end {theorem}
\begin {proof}
Let $\pi $ be an involution of the extended Dynkin diagram. Denote also by $\pi $ the corresponding involution of the set $\{0, 1,\ldots, N\} $ and the induced involution of $ (\t ^\mathbb{ C}) ^* $ which preserves the root system $\Delta$ and satisfies $\pi (\alpha_j) =\alpha_{\pi (j)} $.
Let $\{H_\alpha, R_\alpha\mid\alpha\in\Delta\}$ be a Chevalley basis. That is, writing $\alpha ^\# $ for the dual of the root $\alpha $ with respect to the Killing form $\kappa$ we set $ H_\alpha = (2/ \kappa(\alpha ^\#,\alpha ^\#)){\alpha ^\#} $ and we choose the root vectors $ R_\alpha $ so that
\[
[R_\alpha, R_{-\alpha}] = H_\alpha.
\]
and such that the {\em structure constants} $ c_{\alpha,\beta} $ defined by $ [R_\alpha, R_\beta] = c_{\alpha,\beta} R_{\alpha +\beta} $ satisfy
$ c_{-\alpha, -\beta} = - c_{\alpha,\beta} $.
For any $ b_{\alpha_j}\in\mathbb{ C} $ for $ j = 1,\ldots, N $, we obtain an automorphism $\Theta $ of $\mathfrak{g} ^\mathbb{ C} $ compatible with $\pi $ by requiring that $\Theta (R_{\alpha_j}) = b_{\alpha_j} R_{\pi (\alpha_j)}$ for $ j = 1, \ldots, N $ and that $\{\pi (H_\alpha),\Theta (R_\alpha)\mid\alpha\in\Delta\} $ is a Chevalley basis.
Our first task is to verify that for an appropriate choice of $ b_{\alpha_j} $, the resulting $\Theta $ is an involution.
Given $\pi $ and $ b_{\alpha_1}, \ldots , b_{\alpha_N} $, for any root $\alpha $ we define $ b_{\alpha}\in\mathbb{ C} $ by the equation $\Theta (R_{\alpha}) = b_{\alpha}R_{\pi (\alpha)} $.
The automorphism $\Theta $ will be an involution precisely when $ b_{\alpha_j}b_{\alpha_{\pi (j)}} = 1 $ for $ j = 1, \ldots, N $. For the $j$ with $\pi (j)\neq 0$, we can clearly guarantee this by taking $ b_{\alpha_{\pi (j)}} = b_{\alpha_j}^ {- 1} $.
We will show that $ b_{\alpha_1}, \ldots , b_{\alpha_N} $ can be chosen so that additionally $ b_{\alpha_0}b_{\alpha_{\pi (0)}} = 1 $.
We may express $R_{\alpha_0} $ as
$C[R_{-\beta_1}, [R_{-\beta_2}, \ldots, [R_{-\beta_{K-1}}, R_{-\beta_K}]\ldots]]$ for some non-zero constant $C$ and simple roots $\beta_i$ such that $\sum_{i = 1} ^ K \beta_i = -\alpha_0$.
Now writing $\alpha_0=-\sum_{j = 1} ^ N m_j\alpha_j$ we have
\begin{equation}\label{eq:involbj}
b_{\alpha_0} R_{\alpha_{\pi (0)} =\Theta (R_{\alpha_0}}) = C\prod_{j=1}^N b_{-\alpha_j}^{m_j}[R_{-\pi(\beta_1)}, [R_{-\pi(\beta_2)}, \ldots, [R_{ \pi(\beta_{K-1})}, R_{-\pi(\beta_K)}]\ldots]]
\end{equation}
and $\Theta^2(R_{\alpha_0}) =\prod_{j=1}^N (b_{-\alpha_j}b_{-\alpha_{\pi (j)}})^{m_j} R_{\alpha_0}$, implying
\[
b_{\alpha_0} b_{\alpha_{\pi (0)}} = \prod_{j=1}^N (b_{-\alpha_j}b_{-\alpha_{\pi (j)}})^{m_j}.
\]
Using that $\{\pi (H_{\alpha}),\Theta (R_{\alpha})\mid\alpha\in\Delta\} $ is again a Chevalley basis and that an automorphism of the extended Dynkin diagram must preserve the Killing form gives
\[
b_{\alpha_j}b_{-\alpha_j} =\frac {\kappa (\pi (\alpha_j)) , \pi (\alpha_j)) )} {\kappa (\alpha_j ,\alpha_j )}= 1.
\]
Hence
\[
b_{\alpha_0}b_{\alpha_{\pi (0)}} =\prod_{j = 1} ^ N (b_{\alpha_j}b_{\alpha_{\pi (j)}}) ^ {- m_j} = (b_{\alpha_{\pi (0)}} b_{\alpha_0}) ^ {- 1},
\]
where the last equality uses the assumption $ b_{\alpha_j}b_{\alpha_{\pi (j)}} = 1 $ for $\pi (j)\neq 0 $.
We therefore automatically have $ b_{\alpha_0}b_{\alpha_{\pi (0)}} =\pm 1 $.
Considering \eqref{eq:involbj} shows that if there exists $ j $ such that $\pi (j) = j $ and $ m_j $ is odd then by switching the sign of $ b_{\alpha_j} $ if necessary we may ensure that $ b_{\alpha_0}b_{\alpha_{\pi (0)}} = 1 $.
It remains to give a method of proof for when there is no $\alpha_j$ with $m_j$ odd that is fixed by $\pi $. If $\pi (0) = 0 $ then there is nothing to prove so we assume henceforth that $\pi (0)\neq 0 $. Suppose $\gamma$ is a positive root such that
\begin {enumerate} [(a)]
\item the expression $\gamma =\sum_{j = 1} ^ N n_j\alpha_j $ as a sum of simple roots has $ n_{ \pi(0)} = 0 $,
\item $\pi(\gamma) + \alpha_{\pi(0)}$ is also a root, and
\item $\gamma + \alpha_0 =- \pi(\gamma) - {\alpha_{\pi (0)}} $.
\end {enumerate}
From (c) we have that
\[
[[R_\gamma,R_{\alpha_0}],[R_{\pi(\gamma)}, R_{\pi(0)}]] = c _{\gamma,\alpha_0}c_{\pi (\gamma),\pi (\alpha_0)} H_{\gamma + \alpha_0}.
\]
Applying $\Theta$ gives
\[
[[b_\gamma R_{\pi(\gamma)},b_{\alpha_0}R_{\pi(0)}],[b_{\pi(\gamma)}R_\gamma, b_{\alpha_{\pi (0)}}R_{\alpha_0}]] = - c _{\gamma,\alpha_0}c_{\pi (\gamma),\pi (\alpha_0)} H_{\gamma + \alpha_0}
\]
and so
\begin{align}\label{eq:bgamma}
b_\gamma b_{\pi(\gamma)}b_{\alpha_0}b_{\alpha_{\pi (0)}} = 1.
\end{align}
We may write $R_\gamma$ as $C'[R_{\beta'_1},[R_{\beta'_2} \ldots [R_{\beta'_{K'-1}}, R_{\beta'_{K'}}]]\ldots ]$ with $C'$ a non-zero constant and $\beta'_i\neq\alpha_{\pi(0)}$ simple roots satisfying $\sum_{i = 1} ^ {K'}\beta'_i =\gamma$. Then
\[
b_\gamma b_{\pi (\gamma)} R_\gamma =\Theta^2(R_\gamma) =\left (\prod_{i = 1} ^ {K'} b_{\beta'_i}b_{\beta'_{\pi(i)}}\right) R_\gamma.
\]
However for simple roots $\alpha_j $ with $\pi (j)\neq 0 $ we chose $b_{\alpha_j}$ so that $ b_{\alpha_j} b_{\alpha_{\pi (j)}} = 1 $ and hence
$b_\gamma b_{\pi(\gamma)}=1$. Substituting this into \eqref{eq:bgamma} gives that $b_{\alpha_0}b_{\alpha_{\pi (0)}} = 1$, as required.
A similar argument applies if there are positive roots $\gamma, \delta$ such that
\begin {enumerate} [(i)]
\item the expressions of $\gamma,\delta $ as sums of simple roots do not contain $\alpha_{\pi(0)}$,
\item $\pi(\gamma) + \alpha_{\pi(0)}$ and $\delta + \pi(\delta)$ are also roots, and
\item $\delta +\pi(\delta) + \gamma + \pi(\gamma) = - \alpha_0 - \alpha_{\pi(0)} $.
\end {enumerate}
Here we know there is some non-zero constant $C'' $ such that
$$[[R_\gamma,R_{\alpha_0}],[R_{\pi(\gamma)}, R_{\pi(0)}]] = C'' [R_{-\delta}, R_{-\pi(\delta)}]$$ and as above applying $\Theta$ gives
\[
b_\gamma b_{\pi(\gamma)}b_{\alpha_0}b_{\alpha_{\pi (0)}} = b_{-\delta} b_{-\pi(\delta)}.
\]
By (i)
we know $b_\gamma b_{\pi(\gamma)} = 1$ and $b_{-\delta} b_{-\pi(\delta)}=1$ so conclude that
$b_{\alpha_{\pi (0)}}b_{\alpha_0}=1$.
To show that every involution of the extended Dynkin diagram extends to an involution of the Lie algebra we now consider the involutions of each of the diagrams and, for those that do not fix some $\alpha_j$ with odd $m_j$, identify a suitable root $\gamma$ or pair of roots $\gamma, \delta$.
\setcounter{figure}{0}
\begin{figure}[h]\label{figure:extended}
\begin{tikzpicture}[scale=0.8]
\draw(0,0) node {$E_8$};
\filldraw(1,0) circle (3pt);
\draw(2,0) circle (3pt);
\draw(3,0) circle (3pt);
\draw(4,0) circle (3pt);
\draw(5,0) circle (3pt);
\draw(6,0) circle (3pt);
\draw(7,0) circle (3pt);
\draw(8,0) circle (3pt);
\draw(6,1) circle (3pt);
\draw(1.10,0)--(1.90,0);
\draw(2.10,0)--(2.90,0);
\draw(3.10,0)--(3.90,0);
\draw(4.10,0)--(4.90,0);
\draw(5.10,0)--(5.90,0);
\draw(6.10,0)--(6.90,0);
\draw(7.10,0)--(7.90,0);
\draw(6,0.10)--(6,0.90);
\draw(1,-0.6) node {\footnotesize{$\alpha_0$}};
\draw(2,-0.6) node {\footnotesize{$\alpha_8$}};
\draw(3,-0.6) node {\footnotesize{$\alpha_7$}};
\draw(4,-0.6) node {\footnotesize{$\alpha_6$}};
\draw(5,-0.6) node {\footnotesize{$\alpha_5$}};
\draw(6,-0.6) node {\footnotesize{$\alpha_4$}};
\draw(7,-0.6) node {\footnotesize{$\alpha_3$}};
\draw(8,-0.6) node {\footnotesize{$\alpha_1$}};
\draw(5.5,1) node {\footnotesize{$\alpha_2$}};
\draw(0,2) node {$D_N$};
\draw(1,2) circle (3pt);
\draw(2,2) circle (3pt);
\draw(4,2) circle (3pt);
\draw(5,2) circle (3pt);
\filldraw(2,3) circle (3pt);
\draw(4,3) circle (3pt);
\draw(1.10,2)--(1.90,2);
\draw(2.10,2)--(2.60,2);
\draw(3.40,2)--(3.90,2);
\draw(4.10,2)--(4.90,2);
\draw(3,2) node{$\ldots$};
\draw(2,2.10)--(2,2.90);
\draw(4,2.10)--(4,2.90);
\draw(1,1.6) node {\footnotesize{$\alpha_1$}};
\draw(2,1.6) node {\footnotesize{$\alpha_2$}};
\draw(1.5,3) node {\footnotesize{$\alpha_0$}};
\draw(4,1.6) node {\footnotesize{$\alpha_{N-2}$}};
\draw(3.3,3) node {\footnotesize{$\alpha_{N-1}$}};
\draw(5,1.6) node {\footnotesize{$\alpha_N$}};
\draw(0,4) node {$C_N$};
\filldraw(1,4) circle (3pt);
\draw(2,4) circle (3pt);
\draw(3,4) circle (3pt);
\draw(5,4) circle (3pt);
\draw(6,4) circle (3pt);
\draw[double](1.10,4)--(1.90,4);
\draw(1.80,3.90)--(1.90,4);
\draw(1.80,4.10)--(1.90,4);
\draw(2.10,4)--(2.90,4);
\draw(3.10,4)--(3.60,4);
\draw(4.40,4)--(4.90,4);
\draw(4,4) node {$\ldots$};
\draw[double](5.10,4)--(5.90,4);
\draw(5.20,4.10)--(5.10,4);
\draw(5.20,3.90)--(5.10,4);
\draw(1,3.6) node {\footnotesize{$\alpha_0$}};
\draw(2,3.6) node {\footnotesize{$\alpha_1$}};
\draw(3,3.6) node {\footnotesize{$\alpha_2$}};
\draw(5,3.6) node {\footnotesize{$\alpha_{N-1}$}};
\draw(6,3.6) node {\footnotesize{$\alpha_N$}};
\draw(0,6) node {$B_N$};
\draw(1,6) circle (3pt);
\draw(2,6) circle (3pt);
\draw(4,6) circle (3pt);
\draw(5,6) circle (3pt);
\filldraw(2,7) circle (3pt);
\draw(1.10,6)--(1.90,6);
\draw(2.10,6)--(2.60,6);
\draw(3.40,6)--(3.9,6);
\draw[double](4.1,6)--(4.9,6);
\draw(4.8,6.1)--(4.9,6);
\draw(4.8,5.9)--(4.9,6);
\draw(2,6.10)--(2,6.9);
\draw(3,6)node {$\ldots$};
\draw(1,5.6) node {\footnotesize{$\alpha_1$}};
\draw(2,5.6) node {\footnotesize{$\alpha_2$}};
\draw(4,5.6) node {\footnotesize{$\alpha_{N-1}$}};
\draw(5,5.6) node {\footnotesize{$\alpha_N$}};
\draw(1.5,7) node {\footnotesize{$\alpha_0$}};
\draw(2.658,8.060)arc(-110:-359:1cm);
\draw(4,9)arc(0:-70:1cm);
\draw(3,8)node{$\ldots$};
\draw(0,9)node{$A_N$};
\draw(1.8,9.8)node{\footnotesize{$\alpha_1$}};
\draw(4.3,9.8)node{\footnotesize{$\alpha_N$}};
\draw(4.8,9)node{\footnotesize{$\alpha_{N-1}$}};
\draw(1.6,9)node{\footnotesize{$\alpha_2$}};
\draw(3,10.4)node{\footnotesize{$\alpha_0$}};
\filldraw(3,10)circle(3pt);
\filldraw[color=white](3.707,9.707)circle(3pt);
\filldraw[color=white](4,9)circle(3pt);
\filldraw[color=white](2.293,9.707)circle(3pt);
\filldraw[color=white](2,9)circle(3pt);
\draw(3.707,9.707)circle(3pt);
\draw(4,9)circle(3pt);
\draw(2.293,9.707)circle(3pt);
\draw(2,9)circle(3pt);
\draw(9,8) node {$E_7$};
\draw(10,8) circle (3pt);
\draw(11,8) circle (3pt);
\draw(12,8) circle (3pt);
\draw(13,8) circle (3pt);
\draw(14,8) circle (3pt);
\draw(15,8) circle (3pt);
\filldraw(16,8) circle (3pt);
\draw(10.1,8)--(10.9,8);
\draw(11.1,8)--(11.9,8);
\draw(12.1,8)--(12.9,8);
\draw(13.1,8)--(13.9,8);
\draw(14.1,8)--(14.9,8);
\draw(15.1,8)--(15.9,8);
\draw(13,8.1)--(13,8.9);
\draw(13,9) circle (3pt);
\draw(10,7.6) node {\footnotesize{$\alpha_7$}};
\draw(11,7.6) node {\footnotesize{$\alpha_6$}};
\draw(12,7.6) node {\footnotesize{$\alpha_5$}};
\draw(13,7.6) node {\footnotesize{$\alpha_4$}};
\draw(14,7.6) node {\footnotesize{$\alpha_3$}};
\draw(15,7.6) node {\footnotesize{$\alpha_1$}};
\draw(16,7.6) node {\footnotesize{$\alpha_0$}};
\draw(12.5,9) node {\footnotesize{$\alpha_2$}};
\draw(9,5) node {$E_6$};
\draw(10,5) circle (3pt);
\draw(11,5) circle (3pt);
\draw(12,5) circle (3pt);
\draw(13,5) circle (3pt);
\draw(14,5) circle (3pt);
\draw(12,6) circle (3pt);
\filldraw(12,7) circle (3pt);
\draw(10.1,5)--(10.9,5);
\draw(11.1,5)--(11.9,5);
\draw(12.1,5)--(12.9,5);
\draw(13.1,5)--(13.9,5);
\draw(12,5.1)--(12,5.9);
\draw(12,6.1)--(12,6.9);
\draw(10,4.6) node {\footnotesize{$\alpha_6$}};
\draw(11,4.6) node {\footnotesize{$\alpha_5$}};
\draw(12,4.6) node {\footnotesize{$\alpha_4$}};
\draw(13,4.6) node {\footnotesize{$\alpha_3$}};
\draw(14,4.6) node {\footnotesize{$\alpha_1$}};
\draw(11.5,6) node {\footnotesize{$\alpha_2$}};
\draw(11.5,7) node {\footnotesize{$\alpha_0$}};
\draw(9,3) node {$F_4$};
\filldraw(10,3) circle (3pt);
\draw(11,3) circle (3pt);
\draw(12,3) circle (3pt);
\draw(13,3) circle (3pt);
\draw(14,3) circle (3pt);
\draw(10.1,3)--(10.9,3);
\draw(11.1,3)--(11.9,3);
\draw[double](12.1,3)--(12.9,3);
\draw(12.8,2.9)--(12.9,3);
\draw(12.8,3.1)--(12.9,3);
\draw(13.1,3)--(13.9,3);
\draw(10,2.6) node {\footnotesize{$\alpha_0$}};
\draw(11,2.6) node {\footnotesize{$\alpha_1$}};
\draw(12,2.6) node {\footnotesize{$\alpha_2$}};
\draw(13,2.6) node {\footnotesize{$\alpha_3$}};
\draw(14,2.6) node {\footnotesize{$\alpha_4$}};
\draw(9,1) node {$G_2$};
\filldraw(10,1) circle (3pt);
\draw(11,1) circle (3pt);
\draw(12,1) circle (3pt);
\draw(10.1,1)--(10.9,1);
\draw(11.1,1)--(11.9,1);
\draw(11.1,1.06)--(11.85,1.06);
\draw(11.1,0.94)--(11.85,0.94);
\draw(11.8,0.9)--(11.9,1);
\draw(11.8,1.1)--(11.9,1);
\draw(10,0.6) node {\footnotesize{$\alpha_0$}};
\draw(11,0.6) node {\footnotesize{$\alpha_1$}};
\draw(12,0.6) node {\footnotesize{$\alpha_2$}};
\end{tikzpicture}
\caption{Extended Dynkin diagrams, with the lowest root $\alpha_0 $ coloured.}
\end{figure}
For a root system of type $ A_N $, the simple root coefficients $ m_j = 1 $ for all $ j = 1, \ldots , N $. Thus any diagram involution fixing some node is induced by an involution of the Lie algebra. By inspection of the extended Dynkin diagram shown in Figure~\ref {figure:extended},
we see that when $ N $ is even,
every involution of the extended Dynkin diagram fixes some $\alpha_j $.
When $ N $ is odd we need to consider the rotation $\pi (j) = j + \frac {1} {2} (N + 1)\bmod(N +1) $ and reflections.
For the involution $\pi (j) = j + \frac {1} {2} (N + 1)\bmod(N +1) $, the root $\gamma=\alpha_1 + \alpha_2 +\ldots + \alpha_{\frac12(N-1)}$ satisfies conditions (a), (b), (c) above.
Consider now an involution $\pi $ coming from a reflection. Since we have automatically covered the cases when there is a fixed root we can assume that there are an even number of roots between $\alpha_0$ and $\pi(\alpha_0)$ going in each direction around the circle. Indeed the axis of reflection is between the nodes $(\pi(0)-1)/2$ and $(\pi(0)+1)/2$ and between $(N+\pi(0))/2$ and $(N+\pi(0))/2 + 1$. The roots
$$\gamma = \alpha_1 + \alpha_2 + \ldots + \alpha_{(\pi(0)-1)/2}\quad \text{and} \quad \delta = \alpha_{\pi(0)+1} + \alpha_{\pi(0)+ 2} \ldots +\alpha_{(\pi(0)+N)/2}$$
satisfy conditions (i), (ii), (iii) above.
There is only one involution of the root system of type $B_N$, which sends $\alpha_0$ to $\alpha_1$ and fixes everything else. We can choose $\gamma = \alpha_2 + \ldots + \alpha_N$.
For root systems of type $C_N$ there is again only one involution; $\pi(\alpha_i) = \alpha_{N-i}$. Here choose $\gamma= \alpha_1+\ldots +\alpha_{N-1}$.
For $ D_N $, $ m_1 = m_{N -1} = m_N = 1 $, and so we need only consider involutions which do not fix any of these vertices, of which there are three. These are involutions with $\pi(0) = 1, N-1$ or $N$. If $\pi(0)=1$ then let $\gamma = \alpha_2 + \ldots +\alpha_{N-1}$, and if $\pi(0)=N-1$ or $N$, take $\gamma =\alpha_1+\alpha_2 + \ldots + \alpha_{N-2 }$.
For the root system $ E_6 $, all involutions of the diagram fix the vertex $\alpha_4 $ and $ m_4 = 3 $ is odd.
The unique involution of the extended Dynkin diagram for $ E_7 $ satisfies $\pi (\alpha_0) =\alpha_7 $. A list of all positive roots of $ E_7 $ are tabulated for example in \cite [p 1524-1530] {Vavilov:01}.
Let $\gamma=\alpha_1 + \alpha_2 + 2\alpha_3 + 2\alpha_4 + \alpha_5 + \alpha_6$, so $\pi(\gamma)=\alpha_1 + \alpha_2 + \alpha_3 + 2\alpha_4 + 2\alpha_5 +\alpha_6$ and $\pi(\gamma)+\alpha_{\pi(0)}=\alpha_1 + \alpha_2 + \alpha_3 + 2\alpha_4 + 2\alpha_5 +\alpha_6+\alpha_7$ is also a root. Furthermore $\gamma + \alpha_{\pi(0)} + \pi(\gamma) = 2\alpha_1 +2\alpha_2+ 3\alpha_3 + 4\alpha_4 + 3\alpha_5 +2\alpha_6 +\alpha_7$ which is the highest root.
The extended Dynkin diagrams of type $ E_8, F_4, G_2 $ do not possess any involutions.
We have then shown that given any involution $\pi $ of an extended Dynkin diagram for $ (\mathfrak{g} ^\mathbb{ C},\t ^\mathbb{ C} )$, there exists an involution $\Theta $ of $\mathfrak{g} ^\mathbb{ C} $ preserving $\t ^\mathbb{ C} $ and inducing $\pi $. It remains to show that there is a real form $\mathfrak{g}$ of $\mathfrak{g} ^\mathbb{ C} $ for which $\Theta $ is a Cartan involution and such that $\mathfrak{g}\cap\t ^\mathbb{ C} $ has full rank. For any choice of simple roots we may consider the corresponding {\em Borel subalgebra} $\mathfrak b ^\mathbb{ C} = \mathfrak t ^\mathbb{ C} \oplus \bigoplus_{\alpha\in\Delta^ +} \mathcal G ^\alpha $ and it is easy to see that $\Theta $ preserves the set of simple roots if and only if it preserves the corresponding Cartan and Borel subalgebras. Now by \cite [Theorem 8.6] {Kac:90}
there exists an automorphism $\Psi $ of $\mathfrak{g} ^\mathbb{ C} $ such that $\Psi\Theta\Psi^ {- 1} $ acts on the corresponding simple and lowest root vectors $R_{\alpha_j} $ in the Chevalley basis simply by scaling them by $\pm 1 $, and hence preserves the Cartan and Borel subalgebras $\t ^\mathbb{ C} $ and $\mathfrak b ^\mathbb{ C} $. Then $\Theta $ preserves the Cartan subalgebra $\Psi^ {- 1} (\t ^\mathbb{ C}) $ and the Borel subalgebra $\Psi^ {- 1} (\mathfrak b ^\mathbb{ C}) $ and hence the set of simple roots $\Psi^ {- 1}\{\alpha_1, \ldots ,\alpha_N\} $.
Then there exists a real form $\mathfrak{g}' $ of $\mathfrak{g} ^\mathbb{ C} $ with respect to which $\Theta $ is a Cartan involution and such that $\mathfrak{g}'\Psi^{-1}(\t ^\mathbb{ C}) $ is a Cartan subalgebra of $\mathfrak{g}' $ \cite[proof of Theorem 6.88]{Knapp:02}.
Let $\mathfrak{l} ^\mathbb{ C} $ denote the $ (+1)$-eigenspace of $\Theta $ and $ L ^\mathbb{ C} $ a complex Lie group with Lie algebra $\mathfrak{l} ^\mathbb{ C} $. In \cite [Theorem 1] {Matsuki:79} (c.f. \cite [Proposition 2.1] {Vogan:83}) it was shown that for a given real form $\mathfrak{g}' $ and $\Theta $-stable Cartan subalgebra $\t ^\mathbb{ C} $ of a simple complex Lie algebra $\mathfrak{g} ^\mathbb{ C} $, there exists a $\Theta $-stable Cartan subalgebra $\t' $ of $\mathfrak{g}' $ and $ l\in L ^\mathbb{ C} $ such that $\t ^\mathbb{ C} =\operatorname{Ad}_l (\t') ^\mathbb{ C}$. Hence $\mathfrak{g} =\operatorname{Ad}_l\mathfrak{g}' $ is a real form of $\mathfrak{g} ^\mathbb{ C} $ for which $\t =\mathfrak{g}\cap\t ^\mathbb{ C} $ is a $\Theta $-stable real form of $\t ^\mathbb{ C} $ and $\Theta $ is a Cartan involution of $\mathfrak{g} $.
By Proposition ~\ref {prop:Coxeter} the Coxeter automorphism corresponding to the choice of simple roots $\alpha_1,\ldots,\alpha_N $ preserves the real form $\mathfrak{g} $ and in particular the Cartan subalgebra $\mathfrak t $.\end {proof}
\section{Toda frame}\label{Toda}
We now explore the relationship between cyclic primitive maps and the affine Toda field equations. Henceforth $ G $ shall denote a simple real Lie group, $ T $ a Cartan subgroup and $\alpha_1, \ldots ,\alpha_N $ simple roots such that the resulting Coxeter automorphism $\sigma $ preserves the real group $ G $. This Coxeter automorphism then gives $ G/T $ the structure of a $ k $-symmetric space, where $ k -1 $ is the maximum height of a root of $\mathfrak{g} ^\mathbb{ C} $. We shall consider cyclic primitive maps $\psi $ from the complex plane into $ G/T $ and will see that
cyclic primitive maps $\psi:\mathbb{ C}\rightarrow G/T $
arise from and give rise to solutions of the
two-dimensional affine Toda field equations for $\mathfrak{g} $. Our results also apply to maps from a simply-connected coordinate neighbourhood of any Riemann surface.
The
famous Toda equations arose originally as a model for particle interactions
within a one-dimensional crystal, with the affine model corresponding to the
particles being arranged in a circle. They have been the subject of extensive
study, both as a completely integrable Hamiltonian system and in the context
of Toda field theories.
The standard form of the affine Toda field equation for $\mathfrak{g} $ on the complex plane is
\begin{equation}\label{eq:Toda}
2\Omega_{z{\bar{z}}} = \sum_{j = 0}^ N m_j e^{2 \alpha_j (\Omega)}\alpha_j ^\sharp
\end{equation}
Here $\Omega : \mathbb{ C} \rightarrow i \mathfrak{t}$ is a smooth map, the lowest root $\alpha_0 $ is given by
\[
\alpha_0 = -
\sum_{j = 1}^N m_j \alpha_j,
\]
we set $m_0 = 1$ and $ R_{\alpha_j} $ are root vectors such that $ \alpha_j ^\sharp $ is the dual of $\alpha_j $ with respect to the Killing form. Using \eqref {eq:simplereality}, since the Coxeter automorphism preserves the real form $\mathfrak{g} $ there exists a permutation $ \pi $ of the roots $\alpha_0,\alpha_1,\ldots,\alpha_N $
such that
\begin {equation}\label {eq:permutation}
\overline { \alpha_j} = -\alpha_{\pi (j)}.
\end {equation}
We shall consider the generalisation of the affine Toda field
equations obtained by allowing $m_0$, $m_1, \ldots, m_N $ to be any positive real
numbers such that
$ m_{\pi (j)} =\overline {m_j} $ and $ R_{\alpha_j} $ to be any root vectors satisfying $\overline {R_{\alpha_j}} = R_{-\alpha_{\pi (j)}} $.
Given a cyclic element $W = \sum_{j = 0}^N r_j R_{\alpha_j}$ of
$\mathfrak{g}^{\sigma}_1$, we say that a lift $F : \mathbb{ C}
\to G$ of $\psi : \mathbb{ C} \rightarrow G / T$ is a \emph{Toda frame} with
respect to $W$ if there exists a smooth map $\Omega : \mathbb{ C} \rightarrow i
\mathfrak{t}$ such that
\begin{equation}
\label{eq:Todaframe} F^{- 1} F_z = \Omega_z + \operatorname{Ad}_{\exp \Omega} W.
\end{equation}
We call $\Omega$ an \emph{affine Toda field} with respect to $W$. The
motivation for this nomenclature is
\begin{lemma}
Fix a cyclic element $W = \sum_{j = 0}^N r_j R_{\alpha_j}$ of
$\mathfrak{g}^{\sigma}_1$ such that $ m_{\pi (j)} =\overline {m_j} $ and $\overline {R_{\alpha_j}} = R_{-\alpha_{\pi (j)}} $.
The affine Toda field equation \eqref {eq:Toda} is the integrability condition for the existence
of a Toda frame with respect to $ W $ where we take $m_j = r_j \overline{r_j}$ for $j
= 0, \ldots, N$.
\end{lemma}
\begin{proof}
Using $[R_{\alpha_j}, R_{- \alpha_l}] = 0$ whenever $j \neq l$, we can
rewrite the Toda field equation \eqref {eq:Toda} as
\begin{align*}
2{\Omega}_{z\bar{z}} &
=\sum_{j,l=0}^ Nr_j\overline{r_l}e^{{\alpha}_j({\Omega})}e^{{\alpha}_j({\Omega})}[R_{{\alpha}_j},R_{-{\alpha}_l}]\\
& =[\sum_{j=0}^ N r_je^{{\alpha}_j({\Omega})}R_{{\alpha}_j},
\sum_{l=0}^ N\overline{r_l}e^{{\alpha}_l({\Omega})}R_{-{\alpha}_l}].
\end{align*}
From equation \eqref {eq:roots} we know $e^{\alpha_j (\Omega)} R_{\alpha_j} = \operatorname{Ad}_{\exp
\Omega} R_{\alpha_j}$ and also
\[
e^{\alpha_l (\Omega)} R_{- \alpha_l} = e^{-
\alpha_l (- \Omega)} R_{- \alpha_l} = \operatorname{Ad}_{\exp - \Omega} R_{- \alpha_l} .
\]
If we set $W : = \sum_{j = 0}^ N r_j R_{\alpha_j}$ with the normalisation is described in the lemma then since $\sum_{j = 0} ^ NR_jR_{\alpha_j} =\sum_{j = 0} ^ NR_j\overline {R_{-\alpha_{j}}} $, the Toda field
equation becomes
\[ 2 \Omega_{z \bar{z}} = [ \operatorname{Ad}_{\exp \Omega} W, \operatorname{Ad}_{\exp - \Omega}
\overline{W}]. \]
Now for any given $\Omega : \mathbb{ C} \to i\mathfrak {t} $ the integrability condition for the existence
of a Toda frame with respect to $ W $ is the Maurer-Cartan equation \eqref {eq:MC} for
\[
\varphi = (\Omega_z +
\operatorname{Ad}_{\exp \Omega} W) dz + (\Omega_{\bar z} +
\operatorname{Ad}_{\exp -\Omega} W) d\bar z.
\]
Namely, this integrability condition is
\begin{align*}
0 & =(-{\Omega}_{\bar{z}}+{\operatorname{Ad}}_{\exp
-(\Omega)}{\overline{W}})_z-({\Omega}_z+{\operatorname{Ad}}_{\exp
{\Omega}}W)_{\bar{z}}\\
&\qquad +[{\Omega}_z+{\operatorname{Ad}}_{\exp
{\Omega}}W,-{\Omega}_{\bar{z}}+{\operatorname{Ad}}_{\exp -(\Omega)}{\overline{W}}]\\
& =-2{\Omega}_{z\bar{z}}+[{\operatorname{Ad}}_{\exp {\Omega}}W,{\operatorname{Ad}}_{\exp
-{\Omega}}{\overline{W}}],
\end{align*}
which is precisely the Toda field equation.
\end{proof}
Recall that we write $
\alpha_0 = - \sum_{j = 1}^ N m_j \alpha_j$
for the expression of the lowest root $\alpha_0 $ in terms of the chosen simple roots $\alpha_1,\ldots,\alpha_N $.
Given $\tilde{F} : \mathbb{ C} \to G $ with
\begin {equation}
\label {eq:rootcoefficients}
\tilde F^{-1} \tilde F_z|_{\mathfrak{g}_1 ^\sigma} = \sum_{j=0}^ N c_j R_{\alpha_j},
\end {equation}
we say that a cyclic element
\begin {equation}\label {eq:W}
W=\sum_{j=0}^N r_j R_{\alpha_j}
\end {equation}
$\mathfrak{g} ^\sigma_1 $
is \emph{normalised with respect to} $\tilde{F} : \mathbb{ C} \to G $ if
\[
r_0 \prod_{j = 1}^ N r_j^{m_j} = c_0 \prod_{j = 1}^ N c_j^{m_j}.
\]
\begin{theorem}\label {theorem:Toda}
A map $\psi:\mathbb{ C}\rightarrow G/T $ possesses a Toda frame if and only if it is cyclic primitive
Let $\psi : \mathbb{ C} \to G / T$ be a cyclic primitive map possessing a frame $\tilde{F} : \mathbb{ C} \to G$ such that $c_0 \prod_{j =
1}^ N c_j^{m_j}$ is constant, where $ c_j $ are the root coefficients defined in \eqref {eq:rootcoefficients}.
Then for any cyclic element $W $ of $\mathfrak{g} ^\sigma_1 $ which is normalised with respect to $\tilde F $
there exists a Toda frame $F : \mathbb{ C} \to G$ of $\psi$ with respect to $W$. Furthermore if $\psi$ and $\tilde F$ are doubly periodic with lattice $\Lambda$ then so is the Toda frame $ F $.
Conversely, if $\psi:\mathbb{ C} \to G/T$ has a Toda frame $ F $ with respect to cyclic $ W\in\mathfrak{g}_1 ^\sigma $
then $\psi$ is cyclic primitive and $ W $ is normalised with respect to $ F $. In particular then the root coefficients $c_j$ are such that $c_0 \prod_{j =
1}^ N c_j^{m_j} $ is constant.
\end{theorem}
\begin{proof}
Consider the frames $F : = \tilde{F} \exp X$ of $\psi$ where $X : \mathbb{ C} \to
\mathfrak{t}$. For such $F$ we have $F^{- 1} F_z = \operatorname{Ad}_{\exp - X} \tilde{F} ^{-1}
\tilde{F}_z + X_z$ and so
\[ F^{- 1} F_z |_{\mathfrak{g}_1 ^\sigma} = \operatorname{Ad}_{\exp - X} \tilde{F} ^{-1}
\tilde{F}_z |_{\mathfrak{g}_1 ^\sigma} . \]
This implies the Toda condition of $\operatorname{Ad}_{\exp \Omega} W = F^{- 1} F_z
|_{\mathfrak{g}_1 ^\sigma}$ is equivalent to
\begin{align}
\operatorname{Ad}_{\exp(X+{\Omega})}W=\tilde{F}^{-1}\tilde{F}_z|_{\mathfrak{g}_1 ^\sigma}=\sum_{j=0}^ Nc_jR_{{\alpha}_j}.
\end{align}
Using equation \eqref{eq:roots} we can rewrite this as
\begin{align*}\label{eq:Omega+X}
\sum_{j=0}^ Nr_je^{{\alpha}_j(X+{\Omega})}R_{{\alpha}_j}=\sum_{j=0}^ Nc_jR_{\alpha_j}.
\end{align*}
Comparing root space coefficients implies that
\begin {equation}
\label {eq:Xplus}
e^{\alpha_j (X + \Omega)} =
\frac{c_j}{r_j}\text { for $j = 1, \ldots k$}
\end {equation}
and $r_0 \prod _{j = 1}^ N (e^{\alpha_j
(X + \Omega)})^{- m_j} = c_0$. Since $ W $ is normalised with respect to $\tilde F $
and $\mathbb{ C}$ is simply connected, we can solve for $X + \Omega$. We can then find $\Omega$
and $X$ from $X + \Omega$ by taking its $\mathfrak{t}$ and $i\mathfrak{t}$ components
respectively.
It remains to show that $\Omega_z dz = F ^{-1} \partial F |_{\mathfrak{t}} = \varphi'_{\mathfrak{t}}$.
From the $\mathfrak{g}_1 ^\sigma$ component \eqref {eq:MCg1} of the Maurer-Cartan equation for $\varphi $ we
have
\[ \partial ( \operatorname{Ad}_{\exp \Omega} W) - [ \operatorname{Ad}_{\exp \Omega} W,
\varphi'_{\mathfrak{t}}] = 0 \]
or equivalently
\[ [ \operatorname{Ad}_{\exp \Omega} W, \varphi'_{\mathfrak{t}} - \partial \Omega] = 0. \]
Since $W$ is cyclic so is $\operatorname{Ad}_{\exp \Omega} W$ and thus $\varphi'_{\mathfrak{t}} =
\partial \Omega$.
Conversely, given $W$ and a solution $\Omega$ to the corresponding affine Toda field equation, the resulting Toda frame $F$ is primitive. Furthermore the equation
\[
r_0 (e^{- \sum_{j = 1 } ^ N m_j \alpha_j (X + \Omega)} R_{\alpha_0} + \sum_{j = 1}^ N r_j e^{\alpha_j
(X + \Omega)} R_{\alpha_j} = \sum_{j = 0}^ N c_j R_{\alpha_j}
\]
implies that
$r_0 \prod_{j = 1}^ N r_j^{m_j} = c_0 \prod_{j = 1}^ N c_j^{m_j}$ and hence $c_0 \prod_{j = 1}^ N c_j^{m_j}$ is a non-zero constant. This implies that the $c_j$ are nowhere zero and $\psi$ is cyclic primitive.
Now suppose $\tilde{F}$ is doubly periodic with respect to a lattice $\Lambda $. Then for $j=1,\ldots N$, from\eqref {eq:Xplus} we see that $e^{\alpha_j(X+\Omega)}$ is doubly periodic with respect to $\Lambda$ and so
\[
\exp(X + \Omega) = \exp(\sum_{j=1}^ N \alpha_j(X+\Omega)\eta_j)
\]
is also.
Given any $\Gamma \in \Lambda$ it follows that
\begin{align}\label{eq:conj1}
\exp(X(z+\Gamma) -X(z)) = \exp(\Omega(z)-\Omega(z+\Gamma)).
\end{align}
Using the conjugation map $\mathfrak{g}^\mathbb{ C} \to \mathfrak{g}^\mathbb{ C}$ which fixes $\mathfrak{g}$, we obtain from \eqref{eq:conj1} that
\begin{align}\label{eq:conj2}
\exp(X(z+\Gamma) -X(z)) = \exp(-\Omega(z)+\Omega(z+\Gamma).
\end{align}
When combined, \eqref{eq:conj1} and \eqref{eq:conj2} imply that $\exp(X(z+\Gamma))=\exp(z))$ for all $z$ and hence $\exp X$ is doubly periodic with lattice $\Lambda$.
Since $\tilde F$ and $\exp X$ are both doubly periodic with lattice $\Lambda$ we know $F =\tilde F \exp X$ is also.
\end{proof}
Our chief interest lies in cyclic primitive $\psi $ which are doubly periodic, as it is these we shall show are of finite type. We henceforth restrict our attention to doubly-periodic maps and denote by $ \C/\Lambda $ any genus one Riemann surface.
Let $W$ be a cyclic element of $\mathfrak{g} _1 ^\sigma $ as before. We say that a frame $F : \C/\Lambda \to G$ of $\psi : \C/\Lambda \rightarrow G / T$ is a Toda frame with respect to $W$ if $F$ is a Toda frame of $\psi$ when both are considered as maps from $\mathbb{ C}$. From the proof of Theorem~\ref {theorem:Toda} we make the following observation, which will prove useful in the next section.
\begin{lemma}\label {lemma:periodic}
If $F:\C/\Lambda \to G$ is a Toda frame of $\psi: \C/\Lambda \to G/T$ then the corresponding affine Toda field $\Omega : \mathbb{ C} \rightarrow i\mathfrak{t}$ has the property that $\exp \Omega$ and $\Omega_z$ are doubly periodic with lattice $\Lambda$.
\end{lemma}
\section{Finite type result}\label{finite}
We will now show that all smooth maps $ \psi $ from a 2-torus $\C/\Lambda$ into the $ k $-symmetric space $G/T$ which have a Toda frame are of finite type. Hence all such maps can be constructed from a pair of commuting ordinary differential equations on a finite-dimensional loop algebra. In \cite {BFPP:93} it was shown that all semisimple adapted harmonic maps of a 2-torus into a compact semisimple Lie group are of finite type. We prove our finite type result by adapting the methods of that paper. Note that the existence of a Toda frame forces $\psi $ to be cyclic primitive.
A map $Y : \C/\Lambda \to \mathfrak{g}^\mathbb{ C}$ is called a \emph{Jacobi field} if there exists $\dot\Omega: \C/\Lambda \to \mathfrak{t}^\mathbb{ C}$ such that
\begin{align}\label{eq:Jacobi}
dY + [F^{-1} dF, Y] =
\left(\dot{\Omega}_z + [\dot\Omega, F^{-1} F_z ]\right)dz
+\left(-\dot\Omega_{{\bar{z}}} - [\dot\Omega,F^{-1} F_{{\bar{z}}}]\right)d{\bar{z}}.
\end{align}
If $F_t$ is a family of Toda frames with corresponding $\Omega_t:\mathbb{ C} \to i\mathfrak{t}$ then $\frac{d}{dt} F_t |_{t=0}$ is a Jacobi field with $\dot{\Omega} = \frac{d}{dt}\Omega_t|_{t=0}$. Note that if $\dot{\Omega}=0$ the Jacobi equation is the Killing field equation.
Let $F$ be a Toda frame for $\psi:\mathbb{ C}\rightarrow G/T $. We have
\begin{align*}
F^{-1} dF =
(\Omega_z +\operatorname{Ad}_{\exp \Omega} W )dz + (-\Omega_{{\bar{z}}} + \operatorname{Ad}_{\exp -\Omega}\overline{W})d{\bar{z}}
\end{align*}
for some $\Omega: T ^ 2\to i\mathfrak{t}$ and cyclic $ W\in\mathfrak{g}_1 ^\sigma $. Let $Y$ be a Jacobi field with corresponding $\dot{\Omega}: T ^ 2\to i\mathfrak{t}$.
Then $ Y $ must satisfy
\begin{align}
Y_z + [\Omega_z +\operatorname{Ad}_{\exp \Omega} W, Y]&= \dot{\Omega}_z + [\dot{\Omega},\operatorname{Ad}_{\exp \Omega} W]\label {eq:Jacobi1}\\
Y_{{\bar{z}}} + [-\Omega_{{\bar{z}}} + \operatorname{Ad}_{\exp -\Omega}\overline{W}, Y]&= -\dot{\Omega}_{{\bar{z}}} - [\dot{\Omega},\operatorname{Ad}_{\exp -\Omega}\overline{W}].\label {eq:Jacobi2}
\end{align}
Taking \eqref{eq:Jacobi1}$_{{\bar{z}}} - $ \eqref{eq:Jacobi2}$_z$ we obtain
\begin{equation*}
2\dot{\Omega}_{z\bar{z}}=-\bigl[\operatorname{Ad}_{\exp \Omega} W,[\dot{\Omega},\operatorname{Ad}_{\exp -\Omega}\overline{W}]\bigr] -\bigl[\operatorname{Ad}_{\exp -\Omega}\overline{W},[\dot{\Omega},\operatorname{Ad}_{\exp \Omega} W]\bigr].
\end{equation*}
Since $\Omega$ and $W$ are fixed, we see that $\dot{\Omega}$ satisfies a linear elliptic partial differential equation. As the torus is compact, the space of possible $\dot{\Omega}$ is finite dimensional.
\begin{lemma}\label{lem:Jacobi}
Suppose $\psi: \C/\Lambda\rightarrow G/T$ is a cyclic primitive map possessing
a formal Killing field $Y=\sum_{j\leq 1} \lambda^j Y_j\in \Omega^{\sigma}\mathfrak{g} ^\mathbb{ C}$. Then $\psi$ has a (real) polynomial Killing field with highest term $Y_1$.
\end{lemma}
\begin{proof}
We will find an infinite number of linearly independent Jacobi fields for which some linear combination must be a formal Killing field.
Since $Y$ is a formal Killing field, we have \eqref{eq:lax}.
\[
\sum_{j\leq 1} \lambda^j dY_j
= \left[\sum_{j\leq 1} \lambda^j Y_j,\varphi_\lambda \right].
\]
Comparing coefficients of $\lambda^j$ gives the equations
\begin{align*}
(Y_{j})_zdz + [\varphi'_{\mathfrak{t}} , Y_j] + [\varphi'_{\mathfrak{p}}, Y_{j-1}] &= 0, \\
(Y_{j})_{{\bar{z}}}d\bar{z} + [\varphi''_{\mathfrak{t}}, Y_j] + [\varphi''_{\mathfrak{p}}, Y_{j+1}] &= 0.
\end{align*}
For each $l\in \mathbb{ Z}^+$ set
\[
Y^{l} := \frac12 Y_{-kl} + \sum_{-kl <j \leq 1} \lambda^{j+kl} Y_j.
\]
We will show that the $Y^l$ are all Jacobi fields. Considering the coefficients separately gives
\begin{align*}
(Y^l)_zdz + [ \lambda \varphi'_{\mathfrak{p}} + \varphi'_{\mathfrak{t}} , Y^l]
&= \frac12 (Y_{-kl})_{z}dz + \left[\frac12 Y_{-kl}, \lambda \varphi'_{\mathfrak{p}}\right]\\
(Y^l)_{\bar{z}} d\bar{z}+ \left[\varphi''_{\mathfrak{t}} + \lambda^{-1} \varphi''_{\mathfrak{p}}, Y^l\right]
&= -\frac12 (Y_{-kl})_{{\bar{z}}}d\bar{z} - \left[\frac12 Y_{-kl}, \lambda^{-1} \varphi''_{\mathfrak{p}}\right].
\end{align*}
Since $Y_{-kl} \in \mathfrak{g}_0 = \mathfrak{t}^\mathbb{ C}$ we can set $\dot\Omega^l := \frac12 Y_{-kl}$. With this choice of $\dot{\Omega}$, $Y$ is a solution to \eqref{eq:Jacobi} and hence is a Jacobi field. The space of potential $\dot\Omega$ is finite dimensional, so there must be a non-trivial finite linear combination of the $\dot\Omega^l$ which equals $0$. The corresponding finite linear combination of the $Y^l$ is a formal Killing field. Since the highest order terms of the $Y^{l}$ are each $Y_1$ we can rescale this formal Killing field to one with highest order term $Y_1$.
After multiplying by an appropriate power of $\lambda^k$ we may also assume that the degree of the lowest term has smaller absolute value than the degree of the highest term.
Then $ \overline{\xi} +\xi $
is a polynomial Killing field for $\xi$ and by construction has highest order term $ Y_1 $.
\end{proof}
\begin{theorem}
\label{thm:finite}
Suppose $\psi: \C/\Lambda\rightarrow G/T$ has a Toda frame $ F:\C/\Lambda \to G$. Then $\psi $ is of finite type.
\end{theorem}
\begin{proof}
Let $F:\C/\Lambda \to G$ be the Toda frame of $\psi$ with corresponding $\Omega: \C/\Lambda \to i\mathfrak{t}$ and $ W\in\mathfrak{g}_1 ^\sigma $.
Recall that $\psi$ is of finite type if it has an adapted polynomial Killing field $\xi$, that is a $\xi= \sum_{j=-d}^d \lambda^j \xi_j $ in the real twisted loop algebra $\Omega^\sigma \mathfrak{g}$ satisfying the Killing field equation \eqref{eq:lax} and such that
\[
\xi_d + \lambda\frac 12 \operatorname{Ad}_{\exp \Omega}W.
\]
Since $ G $ was assumed simple, the complexified Lie algebra $\mathfrak{g} ^\mathbb{ C} $ is simple and hence has a faithful linear representation so can be regarded as a subalgebra of some $\mathfrak{gl} (m,\mathbb{ C}) $. If we set
\[
D = d - \operatorname{ad}_{\Omega_z dz- \Omega_{\bar{z}}d\bar{z}}
\]
then we can rewrite \eqref{eq:lax} as
\[
D\xi_\lambda = [\xi_\lambda, (2\Omega_z + \lambda \operatorname{Ad}_{\exp \Omega} W)dz + (-2\Omega_{\bar{z}} + \lambda^{-1}\operatorname{Ad}_{\exp -\Omega} \overline{W})d\bar{z}].
\]
From $d(\operatorname{Ad}_{\exp \Omega} W) = [\Omega_z, \operatorname{Ad}_{\exp \Omega} W]dz +[\operatorname{Ad}_{\exp \Omega} W, \Omega_{\bar{z}}]d\bar{z} $ we know $D\operatorname{Ad}_{\exp \Omega} W =0$.
Writing $V=\ker \operatorname{ad}_{\operatorname{Ad}_{\exp \Omega}W}$ and $V^{\perp} = \im \operatorname{ad}_{\operatorname{Ad}_{\exp \Omega}W}$, we have a bundle decomposition $\C/\Lambda \times \mathfrak{g} ^\mathbb{ C} = V \oplus V^{\perp}$. Furthermore
\[
VV \subset V, \quad V^{\perp}V \subset V^{\perp}, \quad V V^{\perp} \subset V^\perp.
\]
Let $X=\sum_{k\leq -1} \lambda^{k}X_k$ where the $X_k$ are sections of $V^\perp$. We seek $X$ such that
\begin {equation}\label{eq:Y}
Y=(1+X)^{-1}\operatorname{Ad}_{\exp \Omega}W(1+X)
\end {equation}
is a solution of the Killing field equation. Note that
\begin{align*}
DY=(1+X)^{-1}[\operatorname{Ad}_{\exp \Omega}W, DX(1+X)^{-1}](1+X),
\end{align*}
and define a one-form $\kappa$ by
$$\kappa =(1+X)((2\Omega_z + \lambda \operatorname{Ad}_{\exp \Omega}W)dz + (-2\Omega_{\bar{z}} + \lambda^{-1}\operatorname{Ad}_{\exp -\Omega}\overline{W})d\bar{z}-DX)(1+X)^ {- 1}.$$
Routine calculations show that
\begin{align*}
DY + [(2\Omega_z + \lambda \operatorname{Ad}_{\exp \Omega}W)dz &+ (-2\Omega_{\bar{z}} + \lambda^{-1}\operatorname{Ad}_{\exp -\Omega}\overline{W})d\bar{z}, Y]\\
& = (1+X)^{-1} [\operatorname{Ad}_{\exp _\Omega} W, -\kappa](1+ X)
\end{align*}
and hence $Y$ satisfies the Killing field equation if and only if $\kappa$ takes values in $V$.
Our task then is to construct $X$ so that $\kappa$ takes values in $V$. We have
\[
\kappa'\cdot (1+X) = (1+X)(2\Omega_z + \lambda \operatorname{Ad}_{\exp \Omega}W) d z-\partial X
\]
where $\kappa'\cdot (1+ X) $ denotes multiplication.
Note that $\Omega_z $ is valued in $ V^\perp$ as it lies in $\t^\mathbb{ C}$.
The splitting of $\kappa' \cdot (1+X)$ into its $V$ and $V^\perp$ components is
\begin{align*}
(V): & \quad \kappa' = \left (\lambda \operatorname{Ad}_{\exp \Omega}W + (2X\Omega_z)^V\right)dz\\
(V^\perp):& \quad \kappa' \cdot X =\left (2\Omega_z +(2X\Omega_z)^\perp + \lambda X\operatorname{Ad}_{\exp \Omega}W\right)dz -D'X.
\end{align*}
Substitution implies
\begin{align}\label{eq:Vcomp}
\lambda[\operatorname{Ad}_{\exp \Omega}W,X] dz = 2 \Bigl(\Omega_z +(X\Omega_z)^\perp -(X\Omega_z)^V X\Bigr) dz -D' X.
\end{align}
Conversely if \eqref{eq:Vcomp} holds then $\kappa' =\left ( \lambda \operatorname{Ad}_{\exp \Omega}W + (2X\Omega_z)^V\right) dz $ and so $\kappa' $ takes values in $V$.
Comparing the $\lambda^j$ coefficients on both sides of \eqref{eq:Vcomp} we can solve for $X$ inductively over $j$ by at each stage requiring $X_j \in \im \operatorname{ad}_{\operatorname{Ad}_{\exp \Omega}W}$ and
\begin{align*}
[\operatorname{Ad}_{\exp \Omega}W,X_1] &= 2\Omega_z\\
[\operatorname{Ad}_{\exp \Omega}W, X_{j-1}] dz &= 2 \Bigl ((X_j \Omega_z)^\perp -\sum_{s+l=j}(X_s\Omega_z)^V X_l\Bigr) dz -D X_k.
\end{align*}
Define $\nabla_\lambda = d + \operatorname{ad}_{\varphi_\lambda} $ and note that \eqref{eq:flat} says precisely that $\nabla_\lambda$ is a flat connection in the trivial bundle $\C/\Lambda\times\mathfrak{g} ^\mathbb{ C} $.
With $ X $ as above we have $\nabla_\lambda' Y=0$. We wish to show that $\nabla_\lambda'' Y=0$ also, as this will imply that $Y$ satisfies the Killing field equation \eqref {eq:lax}.
Define $B$ by
\begin{equation}\label{eq:B}
\nabla_\lambda'' Y = (1+X)^{-1}B (1+X).
\end{equation}
Using $\operatorname{Ad}_{\exp \Omega}W = (1+X) Y(1+X)^{-1}$ and
\[
\nabla_\lambda'' \operatorname{Ad}_{\exp \Omega}W = [-\Omega_{\bar{z}} + \lambda^{-1} \operatorname{Ad}_{\exp -\Omega}\overline{W}, \operatorname{Ad}_{\exp \Omega}W]d\bar z
\]
we obtain
\begin{align*}
B d\bar z &=[(-\Omega_{\bar{z}} + \lambda^{-1}\operatorname{Ad}_{\exp -\Omega}\overline{W}) d\bar z - \nabla_\lambda'' X (1+X)^{-1},\operatorname{Ad}_{\exp \Omega}W ]
\end{align*}
which shows that $B$ takes values in $V^\perp$.
As $\nabla_\lambda$ is a flat connection we have commutativity of covariant derivatives and hence $\nabla_\lambda'\nabla_\lambda''Y=0$ which we write as
\begin{align}\label{eq:DZB1}
-\nabla_\lambda' X (1+X)^{-1}B + \nabla_\lambda' B + B \nabla_\lambda' X (1+X)^{-1} = 0.
\end{align}
Since $\nabla_\lambda' B = D' B + [2\Omega_z +\lambda \operatorname{Ad}_{\exp \Omega}W,B] dz $, we can rewrite \eqref{eq:DZB1} as
\begin{align}\label{eq:DZB2}
D' B&=[\nabla_\lambda' X (1+X)^{-1} + (\lambda \operatorname{Ad}_{\exp \Omega}W-2\Omega_z ) dz, B].
\end{align}
From its defining equation
\eqref {eq:B} we know that $ B $ is of the form $\sum_{j\leq d}\lambda^{j} B_j$. We will show that $ B =0$. Suppose not, then there is some non-zero top coefficient $B_d$. Since $X$ has only negative powers of $\lambda$, the $\lambda^{d+1}$ term in \eqref{eq:DZB2} is
\[
[\operatorname{Ad}_{\exp \Omega}W, B_d].
\]
However we know that $B_d \in V^\perp$ and hence it must be zero.
Thus $\nabla_\lambda'' Y = 0$ and $Y$ satisfies the Killing field equation. From \eqref{eq:Y} we see that $Y$ is of the form $\sum_{j\leq 0}\lambda^{j}Y_j$ and
furthermore $Y_0=\operatorname{Ad}_{\exp \Omega}W$.
We now need to project this $Y$ onto $\Omega^\sigma(\mathfrak{g}^\mathbb{ C})$ to get a solution to the Killing field equation in the correct loop algebra.
Representations of simple Lie algebras are completely reducible and we have identified $\mathfrak{g}^\mathbb{ C}$ with a subalgebra of $\mathfrak{gl}(m,\mathbb{ C})$ so it must have a complementary subspace in $\mathfrak{gl}(m,\mathbb{ C})$ which is invariant under the adjoint action of $\mathfrak{g}^\mathbb{ C}$. This means there exists a projection map $\pi: \Omega (\mathfrak{gl}(m,\mathbb{ C})) \to \Omega(\mathfrak{g} ^\mathbb{ C})$ such that
\[
d \pi(Y) = \pi(d Y) = \pi([Y, \varphi_\lambda]) = [\pi(Y), \varphi_\lambda].\]
Thus we have that $\pi(Y)\in \Omega(\mathfrak{g} ^\mathbb{ C})$ satisfies the Killing field equation. Furthermore $Y_0 = \pi(\operatorname{Ad}_{\exp \Omega}W) =\operatorname{Ad}_{\exp \Omega}W$. Set $\tilde{Y} = \lambda Y = \sum_{j \leq 1}\lambda^j Y_{j-1}$ and note that $\tilde{Y}_1 = Y_0 = \operatorname{Ad}_{\exp \Omega}W$.
We want to project $\tilde Y$ onto $\Omega^\sigma(\mathfrak{g}^\mathbb{ C})$. Consider the map
$$\pi^{\sigma}_j := \frac 1 {k} (\Id + \epsilon^{-j} \sigma^j + \epsilon^{-2j} \sigma^{2j} + \ldots +\epsilon^{-(k-1)j}\sigma^{(k-1)j})$$ where $\epsilon$ is the $k$-th primitive root of unity. This map $\pi_j^\sigma$ projects any element in $\mathfrak{g}^\mathbb{ C}$ to its part in $\mathfrak{g}_j$.
Thus we can define $\pi^\sigma : \Omega(\mathfrak{g} ^\mathbb{ C}) \to \Omega^\sigma (\mathfrak{g}^\mathbb{ C})$ by
\[
\pi^\sigma (\sum_j \lambda^j \xi_j) = \sum_j \lambda^j \pi_j^\sigma (\xi_j).
\]
Then $\tilde{\xi} = \pi^\sigma (\tilde{Y})$ satisfies
\begin{align*}
d\tilde{\xi}=[\tilde{\xi},\Omega_z + \lambda \operatorname{Ad}_{\exp \Omega} W) dz + ( - \Omega_{{\bar{z}}} +\lambda^{-1}\operatorname{Ad}_{\exp -\Omega}\overline{W}) d\bar{z}]
\end{align*}
and $\tilde{\xi}_1=\tilde{Y}_1=\operatorname{Ad}_{\exp \Omega} W$.
Now we may apply Lemma~\ref{lem:Jacobi} to $\tilde{\xi}$ to conclude the existence of a (real) polynomial Killing field $\xi$ whose top term, $\xi_d$, is $\operatorname{Ad}_{\exp \Omega} W$.
The $d-1$ coefficient of $\xi_z =[\xi, \Omega_z + \lambda\operatorname{Ad}_{\exp \Omega}W]$
is
$$(\operatorname{Ad}_{\exp \Omega}W)_z= \left[\operatorname{Ad}_{\exp \Omega}W, \Omega_z\right] + [\xi_{d-1}, \operatorname{Ad}_{\exp \Omega}W] $$
which implies
$$\left[\xi_{d-1}-2\Omega_z,\operatorname{Ad}_{\exp \Omega}W\right]= 0.$$ Since $W$ is a cyclic element and $\xi_{d-1}-2\Omega_z\in \t$ we conclude $\xi_{d-1}-2\Omega_z=0$ and hence $\xi$ satisfies the theorem.
\end{proof}
\bibliographystyle{plain}
\def$'${$'$}
|
\section{Introduction} \label{sec:intro}
The numerical integration scheme of the orbit-averaged Fokker-Planck (FP) equation developed by \citet{c79} has been one of
the most useful tools for simulating the dynamical evolution of globular star clusters.
In addition to two-body relaxation, many physical processes have been incorporated into FP models to achieve
realistic modelling of the globular cluster evolution;
these processes include tidal cutoff, binary heating, disc and bulge shocks, mass loss via stellar evolution, etc.
(see \citealt*{skt08} for a recent example of detailed FP modelling).
In this paper we consider the dynamical evolution of globular clusters in a steady galactic tidal field.
Our main purpose is to investigate what boundary condition can give a better description
of escape of stars from clusters in the tidal field.
This study has been motivated by the studies by \citet{fh00} and \citet{b01}.
\citet{fh00} found that a large fraction of stars with energies above the escape energy (i.e. potential escapers)
take much longer escape time than the dynamical time.
Until their study it had been generally thought that the escape time-scale is of the order of the dynamical time
and that the mass-loss times of the clusters essentially scale with the relaxation time,
which is much longer than the dynamical time.
The findings of \citet{fh00} indicate that this simple scaling may be spoiled by potential escapers
with long escape times.
In fact \citet{b01} performed $N$-body simulations and showed that the mass-loss times (lifetimes) of clusters
do not scale with the relaxation time $t_{\rm rh}$ but scale with $t_{\rm rh}^{3/4}$.
He concluded that the reason is that some of potential escapers are scattered back to lower energies before they
leave the cluster.
More recently \citet{tf05} showed that the dependence on the relaxation time changes with the strength of the tidal filed.
These two studies have revealed that
the behavior of potential escapers greatly influences the rate of mass loss from clusters in the tidal field.
The effects of long escape times and re-scattering of potential escapers have never been considered
in previous FP models in the literature,
but it was assumed that escapers leave a cluster on the dynamical time-scale,
as is described in detail in Section~\ref{sec:FP}.
Since the effect of the galactic tidal field is essentially important to the cluster evolution,
it is necessary to find a way to include the effect into FP models as precisely as possible.
We should mention that
Takahashi \& Portegies Zwart (1998, 2000) compared FP and $N$-body models of star clusters in the tidal field
and found good agreement between these two theoretical models over a wide range of initial conditions.
They showed that the use of anisotropic FP models with the apocentre escape criterion \citep*{tli97}
and the dynamical-time removal of escapers \citep{lo87}
is necessary to obtain such good agreement.
However, note that in their $N$-body models
the tidal force field is not included but the tidal cutoff is applied.
\citet{tpz00} confirmed that the difference between
tidal cutoff and self-consistent tidal field $N$-body models is small
for a particular set of initial conditions,
but did not do systematic investigations on this problem.
In this study we have devised a new scheme to treat escapers in FP models.
The scheme defines a region of potential escapers in phase space and allows them to be scattered again.
Comparing the results of FP models calculated with the new scheme with the results of $N$-body models,
we examine the accuracy of the FP models.
\section{Fokker--Planck models of star clusters in a steady tidal field} \label{sec:FP}
\subsection{Basic assumptions}
The orbit-averaged FP equation is derived under the assumption of spherical symmetry of star clusters \citep{c79}.
Therefore the tidal field, which is not spherically symmetric, cannot be directly incorporated into
orbit-averaged FP models.
In FP models the effect of the tidal field is taken into account
by imposing a tidal cutoff radius $r_{\rm t}$ on the cluster,
which is treated as an isolated system in other respects.
Under these assumptions the distribution function $f$ of stars at time $t$ depends only on
the energy of a star per unit mass, $E$, and the angular momentum per unit mass, $J$.
\subsection{Classical treatments of escapers}
First we summarise classical treatments of escapers used in FP models of previous studies.
\subsubsection{Escape criteria in phase space} \label{sec:esc_crit}
In previous studies,
two kinds of criteria were adopted to define an escape region in $(E,J)$-space:
\begin{enumerate}
\item Energy criterion
\begin{equation}
E > E_{\rm t} \equiv -\frac{GM}{r_{\rm t}} \label{eq:eng-crit},
\end{equation}
\item Apocentre criterion
\begin{equation}
r_{\rm a}(E,J) > r_{\rm t} \label{eq:apc-crit},
\end{equation}
\end{enumerate}
where $M$ is the cluster mass and $r_{\rm a}(E,J)$ is the apocentre radius of a star having energy $E$ and angular momentum $J$.
It is assumed that a star is destined to escape once it enters into the escape region.
The apocentre criterion \citep{tli97} is considered to be more realistic,
at least as long as the tidal field is modelled as a radial cut-off,
and in fact gives better agreement between FP and $N$-body models \citep{tpz98, tpz00}.
For isotropic FP models, where the distribution function does not depend on $J$,
only the energy criterion can be applied (e.g. \citealt{lo87}).
\subsubsection{Removal of escapers}
In previous studies stars in the escape region are assumed to leave the cluster inevitably,
as mentioned above.
It is also assumed that
the time required for this travel is of the order of the dynamical time at the tidal radius.
Considering this travel time, \citet{lo87} applied the following equation
to the distribution function $f$ in the escape region:
\begin{equation}
\frac{\partial f}{\partial t}
= - \nu_{\rm e} f \left[ 1- \left( \frac{E}{E_{\rm t}} \right)^3 \right]^{1/2}
/t_{\rm tid},
\label{eq:lo}
\end{equation}
where $\nu_{\rm e}$ is a dimensionless constant determining the efficiency of escape
(see also \citealt*{lfr91}).
The time-scale $t_{\rm tid}$ is an orbital time-scale at the tidal radius defined by
\begin{equation}
t_{\rm tid}=\frac{2\pi}{\sqrt{(4\pi/3){G \rho_{\rm t}}}},
\end{equation}
where $\rho_{\rm t}$ is the mean mass density within the tidal radius.
Since the dynamical time is generally much smaller than the relaxation time
in globular clusters, we may assume that escapers leave the cluster immediately after they
enter into the escape region, when we are interested only in the evolution on the relaxation time-scale.
This assumption leads to the boundary condition
\begin{equation}
f=0
\end{equation}
on the tidal boundary (e.g. \citealt{cw90}).
\subsection{A new treatment of escapers}\label{sec:nt}
The boundary condition of equation (\ref{eq:lo}) takes account of the fact that
stars satisfying the escape criterion, i.e. potential escapers, need time to actually leave the cluster.
However the effect of re-scattering of potential escapers is not considered there.
Here we propose a new scheme in which the re-scattering effect is taken into account.
First we summarise basic assumptions and equations.
Suppose that the cluster is on a circular orbit,
with radius $R_{\rm G}$ and angular velocity $\omega$, round the centre of a spherical galaxy.
We consider the motion of a star in the rotating coordinate system moving with the cluster;
the origin is at the cluster centre, the $x$-axis points to the galactic centre,
and the $y$-axis is in the cluster orbital plane.
If the cluster and the galaxy are treated as point masses $M$ and $M_{\rm G}$ ($\gg M$)
and the size of the cluster is much smaller than $R_{\rm G}$,
there exists a conserved quantity known as the Jacobi integral given by
\begin{equation}
E_{\rm J}=\frac{v^2}{2}-\frac{GM}{r}-\frac12 \omega^2 (3x^2-z^2) \label{eq:ej}
\end{equation}
(cf. \citealt{s87}, Chapt. 5).
Here $v$ is the velocity of the star measured in the rotating frame,
$r$ is the distance from the star to the cluster centre,
and the angular velocity $\omega$ is given by
\begin{equation}
\omega=\sqrt{\frac{GM_{\rm G}}{R_{\rm G}^3}} .
\end{equation}
The third term on the right-side in equation (\ref{eq:ej}) is a combination of the centrifugal and tidal potentials.
The effective potential is defined as
\begin{equation}
\phi_{\rm eff}(x,y,z)=-\frac{GM}{r}-\frac12 \omega^2 (3x^2-z^2).
\end{equation}
A contour plot of $\phi_{\rm eff}$ is shown, e.g., in Fig. 5.1 of \citet{s87}.
The effective potential has the saddle points at $(\pm x_{\rm e},0,0)$, where
\begin{equation}
x_{\rm e}=\left( \frac{M}{3M_{\rm G}}\right)^{1/3} R_{\rm G}
\end{equation}
and
\begin{equation}
\phi_{\rm eff}(\pm x_{\rm e},0,0) =-\frac32 \frac{GM}{x_{\rm e}}. \label{eq:pxe}
\end{equation}
The equipotential surface passing through these saddle points intersects with the $y$-axis at $y=\pm y_{\rm e}$, where
\begin{equation}
y_{\rm e}=\frac23 x_{\rm e}. \label{eq:ye}
\end{equation}
The necessary condition for escape of a star from the cluster
is given by
\begin{equation}
E_{\rm J} > E_{\rm J,crit} \equiv -\frac32 \frac{GM}{x_{\rm e}}. \label{eq:ejcrit}
\end{equation}
Note that equations (\ref{eq:pxe}), (\ref{eq:ye}), and (\ref{eq:ejcrit}) are valid
for any spherical galactic potential.
\citet{fh00} found that the time-scale for escape of stars with $E_{\rm J}>E_{\rm J,crit}$
varies as
\begin{equation}
t_{\rm e} \propto ( E_{\rm J}-E_{\rm J,crit})^{-2}. \label{eq:fhte}
\end{equation}
With this relation in mind we have devised a new scheme to follow the evolution of
potential escapers.
In this scheme the evolution of the distribution function $f$ for potential escapers is described by
\begin{equation}
\frac{\partial f}{\partial t} =
\left( \frac{\partial f}{\partial t} \right)_{\rm coll}
-\frac{f}{t_{\rm e}(E)}, \label{eq:fp_pe}
\end{equation}
where the first term on the right-side is the FP collision term
and the second term represents mass loss due to escape.
Here the escape time-scale $t_{\rm e}$ is given by
\begin{equation}
\frac{1}{t_{\rm e}(E)} =
\frac{\nu_{\rm e}}{t_{\rm tid}} \left( 1-\frac{E}{E_{\rm crit}} \right)^2, \label{eq:te}
\end{equation}
where $\nu_{\rm e}$ is a dimensionless numerical constant.
It should be noted that energy $E$, not the Jacobi integral $E_{\rm J}$, is used in equations (\ref{eq:fp_pe}) and (\ref{eq:te}).
Energy $E$ does not include the centrifugal and tidal potentials.
Despite this difference, we use the same critical value of energy
\begin{equation}
E_{\rm crit} = -\frac32 \frac{GM}{r_{\rm t}}, \label{eq:ecrit}
\end{equation}
where the tidal radius $r_{\rm t}$ is identified with $x_{\rm e}$.
One might think that
using equation (\ref{eq:te}) with equation (\ref{eq:ecrit})
is too crude an approximation,
but it brings good agreement between FP and $N$-body models as is shown in Section~\ref{sec:results}.
The most important difference between equations (\ref{eq:lo}) and (\ref{eq:fp_pe})
is that the latter includes the collision term.
Thus equation (\ref{eq:fp_pe}) allows potential escapers to be scattered back to lower energies.
The effect of mass loss is included in both equations in a similar way,
though the functional forms of the escape time-scale $t_{\rm e}$ are different.
In this new treatment of the tidal field,
the escape criteria described in Section~\ref{sec:esc_crit}
are modified as follows:
\begin{enumerate}
\item Energy criterion
\begin{equation}
E > E_{\rm crit} = -\frac32 \frac{GM}{r_{\rm t}} \label{eq:eng-crit_pe},
\end{equation}
\item Apocentre criterion
\begin{equation}
r_{\rm a}(E,J) > \frac23 r_{\rm t} \label{eq:apc-crit_pe}.
\end{equation}
\end{enumerate}
Note that $\phi_{\rm eff}(0,\pm 2r_{\rm t}/3,0) =\phi(0,\pm 2r_{\rm t}/3,0)=-3GM/2r_{\rm t}$.
Equation (\ref{eq:fp_pe}) is applied only in the region where an adopted criterion is satisfied.
\subsection{The Fokker-Planck code}
The FP code used in the present study is essentially the same as
that used by \citet{tpz00},
but adopts the new scheme for treating escapers described above.
The code calculates the evolution of the distribution function $f(E,J,t)$.
Unlike \citet{tpz00}, stellar evolution is not considered in the models presented in this paper.
Instead the effect of heating by three-body binaries is considered in the manner described in \citet{t97}.
For all the models presented in the present paper,
201 energy mesh points, 51 angular-momentum mesh points, and 101 radial mesh points are used.
The meshes are constructed as described in \citet{t95}.
When calculating the evolution of multi-mass clusters,
10 discrete mass-components are used to represent a continuous mass function.
Our FP models have two free parameters: one is $\nu_{\rm e}$ in equation~(\ref{eq:te})
and the other is $\gamma$ in the Coulomb logarithm $\ln (\gamma N)$ appearing in the FP collision term.
How the value of $\nu_{\rm e}$ is determined is described in Section~\ref{sec:results}.
We set $\gamma=0.11$ \citep{gh94a} in most of our runs
and $\gamma=0.02$ \citep{gh96} in a part of runs for multi-mass clusters.
\section{Results} \label{sec:results}
\subsection{Comparison with $N$-body models: single-mass clusters}
First we compare FP models with the full tidal field
models of \citet{b01} and additional $N$-body runs performed for this comparison.
All the model clusters are composed of equal-mass stars
and move on circular orbits round a point-mass galaxy.
The initial distribution of stars is given by King models \citep{k66}.
Results are presented in $N$-body units, where the initial total mass and energy of a cluster
are equal to 1 and $-0.25$, respectively, and the gravitational constant $G=1$.
The same units are used throughout this paper.
Here we will refer to FP models with the boundary condition of equation (\ref{eq:fp_pe}) as ``FPf'' models,
which aim to model clusters in a self-consistent {\it full tidal field}.
FP models with equation (\ref{eq:lo}) will be called ``FPd'' models,
where stars beyond the tidal cutoff radius are removed on the {\it dynamical time-scale}.
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f1.eps}
\end{center}
\caption{Evolution of the cluster mass.
The solid lines represent FPf models,
and the dashed lines represent $N$-body models.
The initial models are $W_0=3$ King models with the number of stars $N=1024$, 4096, 16384 and 65536.
}\label{fig:massW3}
\end{figure}
Fig.~\ref{fig:massW3} compares FPf and $N$-body models concerning
the evolution of the total mass of bound stars.
The initial models are $W_0=3$ King models with the number of stars $N=1024$, 4096, 16384 and 65536.
The new treatment of escapers described by equation (\ref{eq:fp_pe})
with the apocentre criterion of equation (\ref{eq:apc-crit_pe}) is employed in the FPf models.
The agreement between the FPf and $N$-body models is good in all the cases.
In fact the value of the parameter $\nu_{\rm e}$ in equation (\ref{eq:fp_pe})
has been determined so that good agreement is obtained by performing test runs with
different values of $\nu_{\rm e}$ as was done by \citet{tpz00}.
We have finally chosen the value of $\nu_{\rm e}=7$.
All the FPf models shown in Fig.~\ref{fig:massW3} are calculated with this value.
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f2.eps}
\end{center}
\caption{Evolution of
the ratio of the mass of potential escapers $M_{\rm pe}$
to the total cluster mass $M$.
The ratio is plotted as a function of the cluster mass at each instance.
}\label{fig:mpeW3}
\end{figure}
Fig.~\ref{fig:mpeW3} shows the evolution of the ratio of the mass of potential escapers $M_{\rm pe}$
to the total cluster mass $M$ for the runs shown in Fig.~\ref{fig:massW3}.
The agreement between the FPf and $N$-body models is fairly good
also in this comparison.
Note that here $M_{\rm pe}$ for the FPf models is defined as the mass of stars with $E>E_{\rm crit}$,
although the apocentre criterion is used in the simulations.
The mass of stars satisfying the apocentre criterion is smaller than
that of stars with $E>E_{\rm crit}$, but shows a similar trend in time variation.
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f3.eps}
\end{center}
\caption{Half-mass time $T_{\rm half}$ as a function of the initial half-mass relaxation time $t_{\rm rh,i}$
for the initial conditions of $W_0=3$ King models.
Two types of FP models, FPf and FPd models (see text), are shown by
the circles and crosses, respectively,
and $N$-body models are shown by the triangles.
The dotted lines represent scalings proportional to $t_{\rm rh,i}$ and $t_{\rm rh,i}^{3/4}$
(they are arbitrarily shifted in a vertical direction).
}\label{fig:thalfW3}
\end{figure}
\begin{table}
\caption{Half-mass times $T_{\rm half}$ given by $N$-body, FPf, and FPd models
for the initial conditions of King models with $W_0=3$.}\label{tab:W3}
\begin{tabular}{@{}rcccc}
\hline
$N$ & $t_{\rm rh,i}$ & $T_{\rm half}$ & $T_{\rm half}$ & $T_{\rm half}$ \\
& & ($N$-body) & (FPf) & (FPd) \\
\hline
128 & $5.13\times 10^0$ & $8.94\times10^1$ & $8.37\times10^1$ & $6.76\times10^1$ \\
256 & $8.13\times 10^0$ & $1.27\times10^2$ & $1.21\times10^2$ & $9.93\times10^1$ \\
512 & $1.35\times 10^1$ & $1.83\times10^2$ & $1.71\times10^2$ & $1.50\times10^2$ \\
1024 & $2.30\times 10^1$ & $2.59\times10^2$ & $2.51\times10^2$ & $2.41\times10^2$ \\
2048 & $4.01\times 10^1$ & $3.73\times10^2$ & $3.73\times10^2$ & $4.00\times10^2$ \\
4096 & $7.11\times 10^1$ & $5.58\times10^2$ & $5.56\times10^2$ & $6.86\times10^2$ \\
8192 & $1.28\times 10^2$ & $8.41\times10^2$ & $8.33\times10^2$ & $1.20\times10^3$ \\
16384 & $2.32\times 10^2$ & $1.18\times10^3$ & $1.26\times10^3$ & $2.16\times10^3$ \\
32768 & $4.24\times 10^2$ & $1.96\times10^3$ & $1.92\times10^3$ & $3.92\times10^3$ \\
65536 & $7.82\times 10^2$ & $3.05\times10^3$ & $2.96\times10^3$ & $7.22\times10^3$ \\
131072 & $1.45 \times 10^3$ & --- & $4.63\times10^3$ & $1.34\times10^4$ \\
262144 & $2.71 \times 10^3$ & --- & $7.36\times10^3$ & $2.50\times10^4$ \\
524288 & $5.07 \times 10^3$ & --- & $1.19\times10^4$ & $4.68\times10^4$ \\
1048576& $9.54 \times 10^3$ & --- & $1.97\times10^4$ & $8.82\times10^4$ \\
2097152& $1.80 \times 10^4$ & --- & $3.33\times10^4$ & $1.67\times10^5$ \\
\hline
\end{tabular}
\end{table}
Fig.~\ref{fig:thalfW3} shows the half-mass time $T_{\rm half}$,
which is the time required for a cluster to lose a half of its initial mass,
as a function of the initial half-mass relaxation time $t_{\rm rh,i}$.
Here the half-mass relaxation time is defined by
\begin{equation}
t_{\rm rh} = 0.138\frac{N^{1/2}r_{\rm h}^{3/2}}
{G^{1/2} m^{1/2} \ln (\gamma N)} \label{eq:trh}
\end{equation}
(\citealt{s87}, Chapt. 2) with $\gamma=0.11$ \citep{gh94a}.
The results are summarised also in Table~\ref{tab:W3}.
The FPf and $N$-body models show good agreement over the whole range of $N$
where the comparison is made.
The scaling $T_{\rm half} \propto t_{\rm rh}^{3/4}$
gives a reasonable fit to the results of these models
as \citet{b01} found.
The results of FPd models are also shown in Fig.~\ref{fig:thalfW3}.
In these models the parameter $\nu_{\rm e}=2.5$ is used for equation~(\ref{eq:lo}) \citep{tpz00}.
The FPd models show clearly a different scaling from the other models;
$T_{\rm half} \propto t_{\rm rh}$ expect for models with very short $t_{\rm rh}$ (i.e. small $N$ ).
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f4.eps}
\end{center}
\caption{Same as Fig.~\ref{fig:thalfW3},
but FPf models with the energy criterion are compared with those with the apocentre criterion and
the $N$-body models.
}\label{fig:thalfW3e}
\end{figure}
In Fig.~\ref{fig:thalfW3e} FPf models with the energy criterion are compared with
those with the apocentre criterion as well as the $N$-body models.
We have set $\nu_{\rm e}=5$ in the energy-criterion models so that their mass evolution
reasonably agrees with that of the $N$-body models for small $N$.
There is no significant difference between the energy-criterion models and the other models
for $t_{\rm rh,i} \la 100$,
but the energy-criterion models tend to lose mass much faster as $t_{\rm rh,i}$ increases.
This indicates that the apocentre criterion is a better escape criterion for FPf models.
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f5.eps}
\end{center}
\caption{Same as Fig.~\ref{fig:thalfW3}, but
FPf models with $N$ up to $2^{30}$ are shown.
The steeper dotted line represents the relation $T_{\rm half} = t_{\rm rh,i}$.
}\label{fig:thalfW3ex}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f6.eps}
\end{center}
\caption{Logarithmic slope $\alpha = d\log T_{\rm half}/d\log t_{\rm rh,i}$
as a function of the initial half-mass relaxation time $t_{\rm rh,i}$
for the models shown in Fig.~\ref{fig:thalfW3ex}.}
\label{fig:thalfW3expw}
\end{figure}
\begin{table}
\caption{Half-mass times $T_{\rm half}$ given by FPf models
for the initial conditions of $W_0=3$ King models with very large $N$.}\label{tab:W3ex}
\begin{tabular}{@{}cccc}
\hline
$N$ & $t_{\rm rh,i}$ & $T_{\rm half}$ & $T_{\rm half}/t_{\rm rh,i}$\\
\hline
$2^{22}\ (\approx 4.19 \times 10^6)$ & $3.41 \times 10^4$ & $5.77 \times 10^4$ & 1.69 \\
$2^{23}\ (\approx 8.39 \times 10^6)$ & $6.47 \times 10^4$ & $1.02 \times 10^5$ & 1.57 \\
$2^{24}\ (\approx 1.68 \times 10^7)$ & $1.23 \times 10^5$ & $1.82 \times 10^5$ & 1.48 \\
$2^{25}\ (\approx 3.36 \times 10^7)$ & $2.35 \times 10^5$ & $3.31 \times 10^5$ & 1.41 \\
$2^{26}\ (\approx 6.71 \times 10^7)$ & $4.50 \times 10^5$ & $6.09 \times 10^5$ & 1.35\\
$2^{27}\ (\approx 1.34 \times 10^8)$ & $8.62 \times 10^5$ & $1.14 \times 10^6$ & 1.32 \\
$2^{28}\ (\approx 2.68 \times 10^8)$ & $1.65 \times 10^6$ & $2.14 \times 10^6$ & 1.30 \\
$2^{29}\ (\approx 5.37 \times 10^8)$ & $3.18 \times 10^6$ & $4.07 \times 10^6$ & 1.28 \\
$2^{30}\ (\approx 1.07 \times 10^9)$ & $6.12 \times 10^6$ & $7.77 \times 10^6$ & 1.27 \\
\hline
\end{tabular}
\end{table}
As stated above,
the results of the FPf models shown in Fig.~\ref{fig:thalfW3} are reasonably well
described by the scaling law $T_{\rm half} \propto t_{\rm rh,i}^{3/4}$.
However we should not expect this scaling continues to hold in the limit of large $N$.
If this scaling continues, the half-mass time measured in the units of the half-mass relaxation time,
$T_{\rm half}/t_{\rm rh,i}$, would go to zero as $N \to \infty$.
This must be impossible because the mass loss is driven by two-body relaxation.
In order to see the scaling of $T_{\rm half}$ in the limit of large $N$,
we have calculated FPf models with very large $N$,
$N=2^{22}\approx 4.19\times 10^6$ to $2^{30}\approx 1.07\times 10^9$,
which are much larger than typical numbers of stars in globular clusters.
The results of these models are shown in Table~\ref{tab:W3ex} and Fig.~\ref{fig:thalfW3ex}.
In Fig.~\ref{fig:thalfW3ex} we see that $T_{\rm half}$ is nearly proportional to $t_{\rm rh,i}$
for very large $N$ clusters, say, for $t_{\rm rh,i} \ga 10^5$ or $N \ga 10^7$.
This trend is more qualitatively shown in Fig.~\ref{fig:thalfW3expw},
where the change in the logarithmic slope,
\begin{equation}
\alpha = \frac{d \log T_{\rm half}}{d \log t_{\rm rh,i}}, \label{eq:alpha}
\end{equation}
is plotted.
The slope $\alpha$ approaches one as $N$ increases.
The ratio $T_{\rm half}/t_{\rm rh,i} \approx 1.3$ for our largest-$N$ models.
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f7.eps}
\end{center}
\caption{Same as Fig.~\ref{fig:thalfW3}, but for the initial conditions of $W_0=5$ King models.
FPf models with the apocentre criterion and $N$-body models are shown.
}\label{fig:thalfW5}
\end{figure}
\begin{table}
\caption{Half-mass times $T_{\rm half}$ given by $N$-body and FPf models
for the initial conditions of King models with $W_0=5$.}\label{tab:W5}
\begin{tabular}{@{}rccc}
\hline
$N$ & $t_{\rm rh,i}$ & $T_{\rm half}$ & $T_{\rm half}$ \\
& & ($N$-body) & (FPf) \\
\hline
1024 & $2.19 \times 10^1$ & $3.89\times10^2$ & $3.92\times10^2$ \\
2048 & $3.82 \times 10^1$ & $5.78\times10^2$ & $6.07\times10^2$ \\
4096 & $6.77 \times 10^1$ & $9.51\times10^2$ & $9.77\times10^2$ \\
8192 & $1.22 \times 10^2$ & $1.51\times10^3$ & $1.61\times10^3$ \\
16384 & $2.21 \times 10^2$ & $2.54\times10^3$ & $2.67\times10^3$ \\
32768 & $4.04 \times 10^2$ & $4.14\times10^3$ & $4.49\times10^3$ \\
65536 & $7.45 \times 10^2$ & --- & $7.62\times10^3$ \\
131072 & $1.38 \times 10^3$ & --- & $1.31\times10^4$ \\
262144 & $2.58 \times 10^3$ & --- & $2.28\times10^4$ \\
524288 & $4.83 \times 10^3$ & --- & $4.03\times10^4$ \\
1048576 & $9.09 \times 10^3$ & --- & $7.20\times10^4$ \\
2097152 & $1.72 \times 10^4$ & --- & $1.30\times10^5$ \\
\hline
\end{tabular}
\end{table}
We have performed simulations also for the initial conditions of $W_0=5$ King models.
The half-mass times of $N$-body and FPf models for $W_0=5$ are summarised in
Table~\ref{tab:W5} and are plotted in Fig.~\ref{fig:thalfW5}.
Here we find good agreement again.
The same parameter $\nu_{\rm e}=7$ is used for both the $W_0=3$ and $W_0=5$ clusters.
In Fig.~\ref{fig:thalfW5} the slope of the $\log t_{\rm rh,i}$--$\log T_{\rm half}$ relation seems to be in between
$3/4$ and 1.
This point is further examined in subsection \ref{ssec:field_strength}.
\subsection{Dependence on the escape-time function} \label{ssec:escape_time}
\citet{b01} argued that the scaling $T_{\rm half} \propto t_{\rm rh}^{3/4}$
can be explained by a steady state solution of a simple model
for the evolution of potential escapers (see equation (12) of his paper).
His model adopts the escape time-scale $t_{\rm e}$ of equation (\ref{eq:fhte}).
If a different function is assumed for $t_{\rm e}$, his model predicts
a different scaling law.
It is shown that the scaling
\begin{equation}
T_{\rm half} \propto t_{\rm rh}^\frac{\beta+1}{\beta+2} \label{eq:thbeta}
\end{equation}
is obtained for $t_{\rm e} \propto (E-E_{\rm crit})^{-\beta}$ (see Appendix A).
It is interesting to see if this prediction is confirmed by the results of our FPf models.
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f8.eps}
\end{center}
\caption{Same as Fig.~\ref{fig:thalfW3}, but FPf models with
different functional forms of $t_{\rm e}(E) \propto (E-E_{\rm crit})^{-\beta} $ ($\beta=1, 2, 3$) are compared.
The dotted lines represent scalings $t_{\rm rh}^{2/3}$, $t_{\rm rh}^{3/4}$ and $t_{\rm rh}^{4/5}$,
which are predicted by the simple steady-solution model
for $\beta=1$, 2 and 3, respectively (see text).
}\label{fig:thalfW3p}
\end{figure}
We have performed FP runs using a generalized form of equation~(\ref{eq:te}),
\begin{equation}
\frac{1}{t_{\rm e}(E)} =
\frac{\nu_{\rm e}}{t_{\rm tid}} \left( 1-\frac{E}{E_{\rm crit}} \right)^\beta, \label{eq:tebeta}
\end{equation}
with $\beta=1$ and 3.
Fig.~\ref{fig:thalfW3p} plots the half-mass time against the initial half-mass relaxation time
for these runs as well as for the standard runs, where King models with $W_0=3$ are used as initial conditions.
The value of $\nu_{\rm e}$ has been adjusted
so that the non-standard models should have roughly the same half-mass times
with those of the standard ones for lower $N$;
$\nu_{\rm e}=7/3$ and $7\times 3$ for $\beta=1$ and 3, respectively.
The results of the FPf models actually depend on $\beta$,
but the degree of the dependence is weaker than predicted by equation~(\ref{eq:thbeta}).
While this equation predicts the slopes 2/3, 3/4 and 4/5 for $\beta=1$, 2 and 3, respectively,
linear least-squares fitting of the data in Fig.~\ref{fig:thalfW3p}
gives the slopes 0.69, 0.72 and 0.75.
When the fitting is done only for $N \ge 16384$, the slopes are 0.75, 0.75 and 0.77.
Thus the scaling law $T_{\rm half} \propto t_{\rm rh}^{3/4}$
is not a bad approximation in all the cases investigated here.
This is not consistent with equation~(\ref{eq:thbeta}).
\subsection{Dependence on the strength of the tidal field} \label{ssec:field_strength}
\citet{tf05} found that the dependence of $T_{\rm half}$ on $t_{\rm rh,i}$ is affected by the strength
of the tidal field and that the logarithmic slope $\alpha$, defined by equation~(\ref{eq:alpha}),
approaches unity as the strength of the tidal field decreases.
In order to confirm their findings,
we have calculated FPf models for the initial conditions where the initial tidal radius $r_{\rm t,i}$
is greater than the King cutoff radius $r_{\rm K}$ (i.e. the radius at which the density drops to zero)
for each value of $W_0$.
On the other hand, all the models presented above are calculated for the initial conditions with
$r_{\rm t,i}=r_{\rm K}$.
Table~\ref{tab:W3rk} lists the half-mass times
for $W_0=3$ King models with $r_{\rm t,i}/r_{\rm K}=1.4$, 2, 4 and 6,
and Fig.~\ref{fig:thalfW3rk} illustrates these results.
In this figure the results for $W_0=3$ and $W_0=5$ King models with $r_{\rm t,i}/r_{\rm K}=1$
are also plotted.
Note that the ratio $r_{\rm K}(W_0=5)/r_{\rm K}(W_0=3) \approx 1.4$.
Fig.~\ref{fig:thalfpw} shows the variation of $\alpha$ with $t_{\rm rh,i}$.
The results shown in Figs.~\ref{fig:thalfW3rk} and \ref{fig:thalfpw} confirm the findings of \citet{tf05}.
The dependence of $T_{\rm half}$ on $t_{\rm rh,i}$ does depend on the strength of the tidal field.
In the limit of $r_{\rm t,i}/r_{\rm K} \to \infty$ and $N \to \infty$,
it is expected that $\alpha \to 1$.
Note that the curve for $W_0=3$ King models with $r_{\rm t,i}/r_{\rm K}=1.4$
lies very close to that for $W_0=5$ King models with $r_{\rm t,i}/r_{\rm K}=1$
in each of Figs.~\ref{fig:thalfW3rk} and \ref{fig:thalfpw}.
This indicates that the mass-loss time-scale
does not depend very much on the initial concentration of the cluster
but is mainly determined by the strength of the tidal field,
as was found by \citet{tf05}.
\begin{table*}
\begin{minipage}{105mm}
\caption{Half-mass times $T_{\rm half}$ given by FPf models
for the initial conditions of King models with $W_0=3$
and $r_{\rm t,i} > r_{\rm K}$.}\label{tab:W3rk}
\begin{tabular}{@{}rccccc}
\hline
$N$ & $T_{\rm half}$ & $T_{\rm half}$ & $T_{\rm half}$ & $T_{\rm half}$ \\
& ($r_{\rm t,i}/r_{\rm K}=1.4$) & ($r_{\rm t,i}/r_{\rm K}=2$) & ($r_{\rm t,i}/r_{\rm K}=4$) & ($r_{\rm t,i}/r_{\rm K}=6$) \\
\hline
128 & $1.62\times10^2$ & $2.71\times10^2$ & $6.86\times10^2$ & $1.11\times10^3$ \\
256 & $2.22\times10^2$ & $3.48\times10^2$ & $7.61\times10^2$ & $1.12\times10^3$ \\
512 & $3.02\times10^2$ & $4.52\times10^2$ & $8.95\times10^2$ & $1.25\times10^3$ \\
1024 & $4.44\times10^2$ & $6.48\times10^2$ & $1.24\times10^3$ & $1.73\times10^3$ \\
2048 & $6.93\times10^2$ & $1.01\times10^3$ & $1.97\times10^3$ & $2.78\times10^3$ \\
4096 & $1.12\times10^3$ & $1.68\times10^3$ & $3.33\times10^3$ & $4.83\times10^3$ \\
8192 & $1.87\times10^3$ & $2.84\times10^3$ & $5.82\times10^3$ & $8.64\times10^3$ \\
16384 & $3.14\times10^3$ & $4.89\times10^3$ & $1.02\times10^4$ & $1.54\times10^4$ \\
32768 & $5.33\times10^3$ & $8.49\times10^3$ & $1.80\times10^4$ & $2.73\times10^4$ \\
65536 & $9.16\times10^3$ & $1.49\times10^4$ & $3.16\times10^4$ & $4.77\times10^4$ \\
131072 & $1.59\times10^4$ & $2.64\times10^4$ & $5.56\times10^4$ & $8.34\times10^4$ \\
262144 & $2.80\times10^4$ & $4.72\times10^4$ & $9.89\times10^4$ & $1.48\times10^5$ \\
524288 & $4.99\times10^4$ & $8.54\times10^4$ & $1.78\times10^5$ & $2.64\times10^5$ \\
1048576 & $8.98\times10^4$ & $1.56\times10^5$ & $3.24 \times 10^5$ & $4.79 \times 10^5$ \\
2097152 & $1.63\times10^5$ & $2.87 \times 10^5$ & $6.02 \times 10^5$ & $8.99 \times 10^5$ \\
\hline
\end{tabular}
\end{minipage}
\end{table*}
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f9.eps}
\end{center}
\caption{Same as Fig.~\ref{fig:thalfW3}, but FPf models for the initial conditions
of King models with $r_{\rm t,i} > r_{\rm K}$ are compared with
the cases of $r_{\rm t,i} = r_{\rm K}$.
}\label{fig:thalfW3rk}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f10.eps}
\end{center}
\caption{Logarithmic slope $\alpha = d\log T_{\rm half}/d\log t_{\rm rh,i}$
as a function of the initial half-mass relaxation time $t_{\rm rh,i}$.
The models are the same as those shown in Fig.~\ref{fig:thalfW3rk}.
}\label{fig:thalfpw}
\end{figure}
\subsection{Comparison with $N$-body models: multi-mass clusters} \label{ssec:multi-mass}
So far we have concentrated on single-mass clusters.
Here we consider the evolution of multi-mass clusters
comparing our FP models with the $N$-body models of \citet{gb08}.
They performed $N$-body simulations of clusters on circular orbits around a point-mass galaxy.
In their simulations the initial mass function (IMF) is given by $dN/dm \propto m^{-2.35}$
with the ratio $m_{\rm max}/m_{\rm min}=30$.
Stellar evolution is not considered in their simulations.
The clusters initially have the density distribution of King models with $W_0=5$.
The ratio of the initial tidal radius to the King radius $r_{\rm t,i}/r_{\rm K}$ is varied from 1 to 8.
The results of the simulations of \citet{gb08} are summarised in their Table~1.
Note that they use different notations from ours: $r_{\rm J}$ is for the tidal (Jacobi) radius
and $r_{\rm t}$ is for the King radius.
\begin{table}
\caption{Half-mass times $T_{\rm half}$ given by $N$-body \citep{gb08} and FPf models
for the initial conditions of multi-mass King models with $W_0=5$ and $r_{\rm t,i}/r_{\rm K}=1$.
Three sets of the parameters $(\gamma, \nu_{\rm e})$ are used for the FPf models.}
\label{tab:mm_gamma}
\begin{tabular}{@{}rcccc}
\hline
$N$ & $T_{\rm half}$ & $T_{\rm half}$ & $T_{\rm half}$ & $T_{\rm half}$ \\
& ($N$-body) & (FPf) & (FPf) & (FPf) \\
& & (0.11, 7) & (0.02, 7) & (0.02, 40) \\
\hline
1024 & $1.14 \times 10^2$ & $1.20 \times 10^2$ & $1.69 \times 10^2$ & $1.17 \times 10^2$ \\
2048 & $1.74 \times 10^2$ & $1.87 \times 10^2$ & $2.54 \times 10^2$ & $1.80 \times 10^2$ \\
4096 & $2.69 \times 10^2$ & $2.86 \times 10^2$ & $3.75 \times 10^2$ & $2.72 \times 10^2$ \\
8192 & $4.35 \times 10^2$ & $4.39 \times 10^2$ & $5.59 \times 10^2$ & $4.18 \times 10^2$ \\
16384 & $6.70 \times 10^2$ & $6.90 \times 10^2$ & $8.57 \times 10^2$ & $6.59 \times 10^2$ \\
32768 & $1.06 \times 10^3$ & $1.12 \times 10^3$ & $1.36 \times 10^3$ & $1.08 \times 10^3$ \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f11.eps}
\end{center}
\caption{
Half-mass time $T_{\rm half}$ as a function of the initial number of stars $N$
for $W_0=5$ King models with the IMF $dN/dm \propto m^{-2.35}$ ($m_{\rm max}/m_{\rm min}=30$).
FPf models with three different sets of the parameters $\gamma$ and $\nu_{\rm e}$
are compared with
the $N$-body models of \citet{gb08}.
}
\label{fig:mm_gamma}
\end{figure}
FPf models are calculated for the same initial conditions as those of \citet{gb08}.
The results for $r_{\rm t,i}/r_{\rm K}=1$
are summarised in Table~\ref{tab:mm_gamma} and Fig.~\ref{fig:mm_gamma}.
There the results of the FPf models with three different sets of parameters $\gamma$ and $\nu_{\rm e}$ are reported.
\citet{gh94a} estimated the best value of $\gamma=0.11$ for single-mass clusters
by comparing $N$-body models with FP and gas models.
Similarly \citet{gh96} obtained $\gamma=0.02$ for multi-mass with
the IMF $dN/dm \propto m^{-2.5} (m_{\rm max}/m_{\rm min}=37.5)$.
We have calculated FPf models for multi-mass clusters using these two values of $\gamma$.
Fig.~\ref{fig:mm_gamma} shows that the parameter set $(\gamma, \nu_{\rm e}) = (0.11,7)$ adopted for single-mass clusters
gives good fit to the $N$-body models also for multi-mass clusters.
On the other hand the parameter set $(\gamma, \nu_{\rm e}) = (0.02,7)$ results in a clear deviation from the $N$-body models.
If we stick to $\gamma=0.02$,
the value of $\nu_{\rm e}$ needs to be increased to about 40 in order to obtain good agreement with the $N$-body models.
We will discuss in more detail what values of the parameters we should choose in the next section.
\begin{table}
\caption{Half-mass times $T_{\rm half}$ given by FPf models
for the initial conditions of multi-mass King models with $W_0=5$ and $r_{\rm t,i}/r_{\rm K}=2$, 4, 8.
The adopted parameter set is $(\gamma, \nu_{\rm e})=(0.11, 7)$.}
\label{tab:mm_rt}
\begin{tabular}{@{}rccc}
\hline
$N$ & $T_{\rm half}$ & $T_{\rm half}$ & $T_{\rm half}$ \\
& ($r_{\rm t,i}/r_{\rm K}=2$) & ($r_{\rm t,i}/r_{\rm K}=4$) & ($r_{\rm t,i}/r_{\rm K}=8$) \\
\hline
1024 & $3.34 \times 10^2$ & $8.03 \times 10^2$ & $1.81 \times 10^3$ \\
2048 & $5.12 \times 10^2$ & $1.17 \times 10^3$ & $2.41 \times 10^3$ \\
4096 & $7.79 \times 10^2$ & $1.72 \times 10^3$ & $3.35 \times 10^3$ \\
8192 & $1.21 \times 10^3$ & $2.65 \times 10^3$ & $5.08 \times 10^3$ \\
16384 & $1.95 \times 10^3$ & $4.30 \times 10^3$ & $8.33 \times 10^3$ \\
32768 & $3.26 \times 10^3$ & $7.32 \times 10^3$ & $1.45 \times 10^4$ \\
\hline
\end{tabular}
\end{table}
\begin{table}
\caption{
Same as Table~\ref{tab:mm_rt}, but
the results of FPf models with the parameter set $(\gamma, \nu_{\rm e})=(0.02, 40)$ are listed.}
\label{tab:mm_rt2}
\begin{tabular}{@{}rccc}
\hline
$N$ & $T_{\rm half}$ & $T_{\rm half}$ & $T_{\rm half}$ \\
& ($r_{\rm t,i}/r_{\rm K}=2$) & ($r_{\rm t,i}/r_{\rm K}=4$) & ($r_{\rm t,i}/r_{\rm K}=8$) \\
\hline
1024 & $3.68 \times 10^2$ & $9.28 \times 10^2$ & $2.20 \times 10^3$ \\
2048 & $5.53 \times 10^2$ & $1.32 \times 10^3$ & $2.87 \times 10^3$ \\
4096 & $8.25 \times 10^2$ & $1.89 \times 10^3$ & $3.86 \times 10^3$ \\
8192 & $1.26 \times 10^3$ & $2.84 \times 10^3$ & $5.63 \times 10^3$ \\
16384 & $2.01 \times 10^3$ & $4.53 \times 10^3$ & $8.92 \times 10^3$ \\
32768 & $3.55 \times 10^3$ & $7.62 \times 10^3$ & $1.52 \times 10^4$ \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[width=84mm]{f12.eps}
\end{center}
\caption{
Same as Fig.~\ref{fig:mm_gamma}, but FPf models
are compared with the $N$-body models of \citet{gb08}
for the clusters with $r_{\rm t,i} > r_{\rm K}$.
The adopted parameter sets for the FPf models are $(\gamma, \nu_{\rm e})=(0.11, 7)$ and $(0.02, 40)$.
}
\label{fig:mm_rt}
\end{figure}
The results for the initial conditions with $r_{\rm t,i}>r_{\rm K}$
are shown in Tables~\ref{tab:mm_rt} and \ref{tab:mm_rt2} and Fig.~\ref{fig:mm_rt}.
The results of \citet{gb08} are not shown in these tables (see their Table 1).
Fig.~\ref{fig:mm_rt} shows that the FPf models with $(\gamma, \nu_{\rm e})=(0.11, 7)$
are in good agreement with the $N$-body models
for $r_{\rm t,i}/r_{\rm K}=2$ and 4.
The FPf models with $(\gamma, \nu_{\rm e})=(0.02, 40)$ are a little farther to the $N$-body models
but still follow them rather well.
However, for $r_{\rm t,i}/r_{\rm K}=8$, a noticeable difference is observed between the FPf and $N$-body models;
in Fig.~\ref{fig:mm_rt} the curve for the $N$-body models is approximately linear
but the slopes of the curves for the FPf models apparently change with $N$.
Neither parameter set reproduces the results of the $N$-body models as well as
in the cases of $r_{\rm t,i}/r_{\rm K}<8$.
The reason for this discrepancy is not clear at present, but
there is a possibility that very early core-collapse in the models with $r_{\rm t,i}/r_{\rm K}=8$ is,
at least partially, responsible for it.
The FPf model with $(\gamma, \nu_{\rm e})=(0.11, 7)$ and $r_{\rm t,i}/r_{\rm K}=8$
experiences core collapse (bounce) at $t=0.006 T_{\rm half}$ for $N=1024$,
and at $t=0.03 T_{\rm half}$ for $N=32768$.
The Coulomb logarithm may take different values for pre-collapse and post-collapse stages
(see the next section), which affects the time-scale of the evolution of FP models.
\section{Discussion}
We have shown that FP models can well follow the mass evolution of star clusters in a tidal field
if a new scheme for treating potential escapers is implemented.
This is the first time the effect of re-scattering of potential escapers has been taken into account
in FP models.
Although Takahashi \& Portegies Zwart (1998, 2000) showed that anisotropic FP models are in good agreement
with $N$-body models for the mass evolution of star clusters in a galaxy,
the tidal field is treated as a tidal cutoff rather than an actual force field.
In the present study we have found that our new FP models are in good agreement
with $N$-body models calculated with the inclusion of the tidal force field.
Thus the new scheme has improved the accuracy of FP models.
\citet{b01} argued that some potential escapers are scattered back to lower energies
before they leave the cluster and that this complicates the scaling of the mass-loss time.
The success of our models is consistent with his argument.
Actually our equation for potential escapers, equation~(\ref{eq:fp_pe}),
can be regarded as a generalization of the equation of his toy model, his equation (12),
used for explaining the scaling $T_{\rm half} \propto t_{\rm rh}^{3/4}$.
The toy model of \citet{b01} is useful
for giving us insight into the effect of potential escapers on the cluster evolution.
On the other hand, the results presented in subsection~\ref{ssec:escape_time}
have revealed the limitation of the model.
When the energy dependence of the escape time is artificially changed from the true one,
the toy model does not correctly explain the results of our FP models.
This failure of the toy model is not a big surprise,
because it is only a simplified model based on many assumptions, some of which are not very realistic.
For example, our simulations show that an exact steady state is never established,
but the toy model assumes a steady state.
In addition, the scaling of the cluster lifetime depends on the strength of the tidal field,
as found by \citet{tf05} and confirmed by the present study,
but the toy model does not take account of the strength of the tidal field.
Our FP models show good agreement with $N$-body models not only for single-mass clusters
but also for multi-mass clusters.
However, we have encountered a difficulty in determining proper values of the two parameters,
$\gamma$ and $\nu_{\rm e}$, in the FP models.
As shown in subsection~\ref{ssec:multi-mass},
the parameter set $(\gamma, \nu_{\rm e})=(0.11, 7)$ brings good agreement
for both single-mass and multi-mass clusters.
Since the escape time-scale $t_{\rm e}$ given by equation~(\ref{eq:te}) is expected to be independent of stellar mass,
it is natural that the same value of the parameter $\nu_{\rm e}$ is applicable
to both single-mass and multi-mass clusters.
On the other hand, the value of $\gamma$ is expected to depend on the stellar mass function.
\citet{h75} argued theoretically that the value of $\gamma$ is generally smaller in multi-mass clusters
than in single-mass clusters.
Based on the results of $N$-body simulations,
\citet{gh94a} obtained a value of $\gamma=0.11$ for isolated single-mass clusters,
and \citet{gh96} obtained a much smaller value, $\gamma=0.02$, for isolated multi-mass clusters
having an IMF similar to the IMF used in our simulations.
When we adopt the value of $\gamma=0.02$ for multi-mass clusters,
we have to use a much larger value of $\nu_{\rm e}$, $\nu_{\rm e}=40$,
than the best value of $\nu_{\rm e}=7$ for single-mass clusters,
in order to obtain good agreement with $N$-body models.
Thus we have not found a parameter set satisfying both
the independence of $\nu_{\rm e}$ on the mass function and
the dependence of $\gamma$ on it.
It needs further investigation to solve this incompatibility,
but even the determination of $\gamma$ itself is not a simple task.
For example, \citet{gh94b} obtained the best value of $\gamma=0.035$
by examining the post-collapse evolution of $N$-body models
of isolated single-mass clusters.
This value is much smaller than the value of $\gamma=0.11$ obtained for
pre-collapse single-mass clusters.
These results suggest that the value of $\gamma$
changes along with the evolution of clusters.
It may also change with radius within a cluster \citep{gh94a}.
\citet{fh00} theoretically estimated not only the energy dependence of the escape time-scale $t_{\rm e}$
but also its numerical coefficient, which is given in their equation (9).
If we ignore the difference between energy $E$ and the Jacobi integral $E_{\rm J}$,
their estimate for a $W_0=3$ King model leads to a value of $\nu_{\rm e}=29$.
This is about four times larger than our best value of $\nu_{\rm e}=7$ for single-mass clusters.
However,
\citet{fh00} also did numerical experiments and found that their theoretical estimate of $t_{\rm e}$ is too small;
escape time-scales obtained from the numerical experiments are more than a few times larger than the theoretical one.
Therefore our value $\nu_{\rm e}=7$ is not inconsistent with the result of \citet{fh00}.
On the other hand,
our value of $\nu_{\rm e}=40$ for multi-mass clusters with $\gamma=0.02$
is a little larger than their theoretical estimate.
Another issue not addressed in the present paper is how the mass profile of the parent galaxy
affects the results.
In all the simulations presented here
we assume that the parent galaxy is represented by a point mass.
On the other hand, \citet{tf10} showed that the mass-loss time-scale depends on
the mass profile of the parent galaxy;
the time-scale increases as the mass profile gets shallower.
Therefore we expect that the parameter $\nu_{\rm e}$ depends on the mass profile of the parent galaxy.
This issue will be examined in a future study.
\section{Conclusion}
In this paper we have developed new FP models of globular clusters in a steady galactic tidal field.
Our FP models are novel in the method of treating escapers:
potential escapers are allowed to experience gravitational scattering with other stars
before they really leave clusters.
The new method has been devised in order to construct more realistic models of star clusters in a tidal field
compared to simple tidal-cutoff models as in previous studies.
The mass evolution of clusters in a tidal field does not simply scale with the relaxation time,
and our FP models are in good agreement with $N$-body models in this respect.
Our FP models include two parameters $\gamma$ and $\nu_{\rm e}$;
$\gamma$ is the numerical factor in the Coulomb logarithm $\ln (\gamma N)$ and
$\nu_{\rm e}$ adjusts the speed of the tidal mass loss.
We have determined the best values of $\nu_{\rm e}$ for given values of $\gamma$
by comparing FP results with $N$-body results.
For single-mass clusters the best parameter set is $(\gamma, \nu_{\rm e})=(0.11, 7)$.
This parameter set is applicable to multi-mass clusters as well,
but another set $(\gamma, \nu_{\rm e})=(0.02, 40)$ does work equally well as long as multi-mass clusters are concerned.
The parameter $\nu_{\rm e}$ is expected to depend on the mass profile of the parent galaxy,
though a point-mass galaxy is assumed in all the simulations of the present paper.
Further investigation is required for the determination of the best values of the parameters
$\gamma$ and $\nu_{\rm e}$ under various conditions.
While FP models are generally thought to be less faithful models of globular clusters than $N$-body models,
the present study has significantly improved the accuracy of FP models.
An advantage of FP models is that they can be calculated much faster than $N$-body models.
Therefore FP models are particularly useful when we need to calculate a huge number of models.
For example, when we try to specify the initial conditions of individual clusters,
we have to perform simulations for many sets of the initial conditions,
because the parameter space to be searched is very large.
We believe that our FP models is quite useful for such searching.
\section*{Acknowledgments}
Part of the work was done while the authors visited the Center for
Planetary Science (CPS) in Kobe, Japan, during a visit that was
funded by the HPCI Strategic Program of MEXT.
We are grateful for their hospitality.
HB acknowledges support by the Australian Research Council (ARC) through Future Fellowship Grant FT0991052.
The numerical calculations of the Fokker-Planck models were carried out
on Altix3700 and SR16000 at YITP in Kyoto University.
|
\section{Introduction}
\label{sec:intro}
\subsection{Collective phenomena in noisy coupled oscillators}
Coupled oscillator models are omnipresent in the scientific literature because
the emergence of coherent behavior in large families of interacting units that have
a periodic behavior, that we generically call {\sl oscillators}, is an extremely common
phenomenon (crickets chirping, fireflies flashing, planets orbiting, neurons firing,...).
It is impossible to properly account for the literature and the various models
proposed for
this kind of phenomena, but while a precise description of each of the different
instances in which synchronization emerges demands specific, possibly very complex,
models, the {\sl Kuramoto
model} has emerged as capturing some of the
fundamental aspects of synchronization \cite{cf:acebron}.
It can be introduced
via
the system of $N$ stochastic differential equations
\begin{equation}
\label{eq:Kmod}
\dd\varphi^\omega_j(t)\, =\,
\omega_j \dd t - \frac KN \sum_{i=1}^N \sin(\varphi^\omega_j(t)-\varphi^\omega_i(t)) \dd t + \sigma \dd B_j(t)\, ,
\end{equation}
for $j=1, \ldots, N$, where
\begin{enumerate}
\item $\{B_j\}_{j=1, \ldots, N}$ is a family of standard independent Brownian motions: in physical terms,
this is a {\sl thermal noise};
\item $\{\omega_j\}_{j=1\cdots N}$ is a family of independent identically distributed random variables of law $\mu$: they are are the {\sl natural
frequencies} of the oscillators and, in physical terms,
they can be viewed as a {\sl quenched disorder};
\item $K$ and $\sigma $ are non-negative parameters, but one should think of them as
positive parameters since the cases in which they vanish have only a marginal role
in the what follows.
\end{enumerate}
\medskip
The variables $\varphi^\omega_j$ are meant to be angles (describing the position of rotators on the circle
${\ensuremath{\mathbb S}} $), so we focus on $\varphi^\omega_j\;\text{mod}\;2\pi$ and
\eqref{eq:Kmod} defines, once an initial condition is supplied, a
diffusion process on ${\ensuremath{\mathbb S}} ^N$. Note that if
$\{ \varphi^\omega _j (\cdot)\}_{j =1, \ldots, N}$ solves \eqref{eq:Kmod}, also
$\{ \varphi^\omega _j (\cdot)+ \varphi\}_{j =1, \ldots, N}$, with $\varphi\in {\ensuremath{\mathbb S}} $, is a solution: this is the
rotation symmetry of the system that will repeatedly make surface
in the remainder of the paper.
\medskip
Some of the main features \eqref{eq:Kmod} are easily grasped:
each oscillator rotates at its own speed, it is perturbed by independent
noise and it interacts with all the other oscillators: the interaction tends to
align the rotators. It may be helpful at this stage to point out
that if $\mu = \delta_0$, that is the natural frequencies are just zero, then the dynamics is reversible with invariant probability measure that, up to normalization, is
\begin{equation}
\label{eq:mfr}
\exp\left(
\frac{K}{\sigma^2}
\sum_{i,j=1}^N \cos \left( \varphi_i -\varphi_j\right)
\right) \lambda_N (\dd \varphi)\, ,
\end{equation}
where $\lambda_N$ is the uniform measure on ${\ensuremath{\mathbb S}} ^N$.
The Gibbs measure in \eqref{eq:mfr} is a well known statistical mechanics model --
it is the classical XY spin mean field model or rotator mean field model --
treated analytically in \cite{cf:SFN,cf:Pearce} in the $N \to \infty$ limit. In particular,
the model exhibits a phase transition at $K=K_c:= 1/\sigma^2$, that is effectively
a {\sl synchronization transition}: in the $N \to \infty$ limit we have that for $K\le K_c$
the rotators become independent and
uniformly distributed over ${\ensuremath{\mathbb S}} $, while for $K>K_c$ the limit measure is obtained by choosing a phase $\theta$ uniformly in ${\ensuremath{\mathbb S}} $ and by choosing the values of the
phase of each oscillator by drawing it at random following a
suitable distribution that concentrates around $\theta$.
However, in \cite[Prop.~1.2]{cf:BGP}, it is shown that, unless $\mu=\delta_0$,
the model is not reversible (for $\mu$ almost surely all the realization of $\omega$)
and one effectively steps into the domain of non-equilibrium statistical mechanics.
Our approach actually relies on a sharp control of the reversible case and works when
the system is not too far from reversibility, that is for weak disorder. Our approach
actually applies well beyond \eqref{eq:Kmod}: here we will treat explicitly
the case $\omega_j$ is replaced by $U(\varphi_j^\omega, \omega_j)$, that is the natural frequency
$\omega_j$ is replaced by a {\sl natural dynamics} that can be substantially different
from one oscillator to another. This model is a disordered version of the active
rotator model considered for example in \cite{cf:shinomoto1986a}.
Since we will focus on $\sigma >0$, from now on, for ease of exposition, we set $\sigma :=1$.
\subsection{The Fokker-Planck or McKean-Vlasov limit}
An efficient way to tackle \eqref{eq:Kmod} is to consider the empirical probability
on ${\ensuremath{\mathbb S}} \times {\ensuremath{\mathbb R}} $
\begin{equation}
\label{eq:empP}
\nu_{N, t}^\omega (\dd \theta, \dd \omega )\, :=\, \frac 1N \sum_{j=1}^N
\delta_{(\varphi_j^\omega(t), \omega_j)} (\dd \theta, \dd \omega)\, .
\end{equation}
In fact, in the $N\to \infty$ limit, the sequence of measures
$\{ \nu_{N, t}^\omega\}_{N=1,2 , \ldots}$ converges to a limit measure whose density (with respect
to $\lambda_1 \otimes \mu$) solves the nonlinear
Fokker-Planck equation
\begin{equation}
\label{FKP kuramoto disorder}
\partial_t p_t(\theta,\omega)\, =\, \frac{1}{2} \Delta p_t(\theta,\omega) -\partial_\theta \Big(p_t(\theta,\omega)(\langle
J*p_t\rangle_\mu(\theta) +\omega)\Big),
\end{equation}
where $J(\theta)=-K\sin(\theta)$, $\ast$ denotes the convolution and $\langle \cdot\rangle_\mu$ is a notation for the integration with respect to $\mu$, so $\langle J\ast u\rangle_\mu(\theta) = \int_{\bbR}\int_{\bbS}{J(\varphi)
u(\theta-\varphi, \omega)\dd\varphi\mu(\dd\omega)}$ is the convolution of $J$ and $u$, averaged with respect to the
disorder.
Here and throughout the whole paper $\Delta$ means $\partial^2 _\vartheta$.
The Fokker-Planck PDE \eqref{FKP kuramoto disorder} appears repeatedly in the physics and biology literature, see e.g. \cite{cf:acebron,cf:Sakaguchi,cf:StrogatzMirollo}, and a mathematical
proof (and precise statement) of the result we just stated can be found
in \cite{cf:dPdH,cf:eric}. Notably, in \cite{cf:eric} the result is established
under the assumption that $\int \vert \omega \vert \mu( \dd \omega)< \infty$ and emphasis
is put on the fact that the result holds for almost every realization of the disorder
sequence $\{ \omega_j\}_{j=1,2, \ldots}$. Let us point out that in \eqref{FKP kuramoto disorder} $\omega$ is a one dimensional real variable, while in \eqref{eq:Kmod} the superscript
$\omega$ is a short for the whole sequence of natural frequencies. Since what follows
is really about \eqref{FKP kuramoto disorder} this abuse of notation will be of limited impact.
In Appendix \ref{sec:appendix regularity pt with disorder}, we detail the fact that
\eqref{FKP kuramoto disorder} generates an evolution
semigroup in suitable spaces. Here we want to stress that
\eqref{FKP kuramoto disorder} can be viewed as a family of coupled
PDEs, one for each value of $\omega$ in the support of $\mu$: $p_t(\cdot, \omega)$
is the distribution of phases in the population of oscillators with natural
frequency $\omega$.
\subsection{About stationary solutions to \eqref{FKP kuramoto disorder}}
Remarkably (\cite{cf:Sakaguchi}, see also \cite{cf:dH}),
if $\mu$ is symmetric all the
stationary solutions to \eqref{FKP kuramoto disorder} can be written in a
semi-explicit way as ${q}(\theta+\theta_0, \omega)$ ($\theta_0$ is an arbitrary constant
that reflects the rotation symmetry) where
\begin{equation}
\label{eq:qhatom}
{q}(\theta, \omega) \,:=\, \frac{S(\theta, \omega, 2Kr)}{Z(\omega,2Kr)}\, ,
\end{equation}
with
\begin{equation}
\label{eq:Sq}
S(\theta, \omega, x) \,=\, e^{G(\theta, \omega, x)}\left[ (1-e^{4\pi\omega})
\int_{0}^{\theta}{e^{-G(u,
\omega, x)}\dd u} + e^{4\pi\omega}\int_{0}^{2\pi}{e^{-G(u, \omega, x)} \dd u} \right],
\end{equation}
and $G(u, y, x)= x\cos(u) + 2y u$, $Z(\omega,x)= \int_{\bbS} S(\theta, \omega, x) \dd\theta$ is the
normalization constant and
$r\in[0,1]$ satisfies the fixed-point relation
\begin{equation}
\label{eq:fixedpointom}
r\,=\, \Psi^\mu(2Kr), \ \ \ \text{where} \ \ \ \Psi^\mu(x)\,:=\, \int_{\bbR}\frac{\int_{\bbS}\cos(\theta)S(\theta, \omega,
x)\dd\theta}{Z(\omega,x)}\mu(\dd\omega)\, .
\end{equation}
A series of remarks are in order:
\begin{enumerate}
\item $r=0$ solves \eqref{eq:fixedpointom} and this corresponds to the
fact that $q(\cdot)\equiv\frac1{2\pi}$ is a stationary solution. It is the only
one as long as $K$ does not exceed
critical value $K_c$ which is in any case not larger than
\begin{equation}
\label{eq:Ktilde}
\widetilde{K} \, :=\,
\left( \int_{\ensuremath{\mathbb R}} \frac{\mu(\dd \omega)}{1+4\omega^2} \right)^{-1}\, ,
\end{equation}
as one can easily see by computing (see e.g. \cite{cf:dH}) the derivative of
$\Psi^\mu(2K\cdot)$ at the origin and noticing that is larger than one if and only if
$K >\tilde K$ and that $\Psi^\mu(\cdot) <1$, see Figure~\ref{fig:fixed point Psi}.
\item When \eqref{eq:fixedpointom} admits a fixed point $r>0$, and this
is certainly the case if $K>\widetilde K$, a nontrivial
stationary solution is present and in fact, by rotation symmetry, a circle
of non-trivial stationary solutions. Such solutions correspond to a synchronization phenomenum, since the distribution of the phases is no longer
trivial.
\item As explained in Figure~\ref{fig:fixed point Psi} and its caption, in general there can be
more than one fixed point $r>0$: in absence of disorder there is only one
positive fixed point (when it exists, that is for $K>1$), but this fact is non-trivial
even in this case (see below). Uniqueness is expected for $\mu$ which is unimodal, but
this has not been established.
\item While the local stability of $\frac 1{2\pi}$ is understood
\cite{cf:StrogatzMirollo} and it holds only if $K \le \widetilde{K}$, the stability properties of the non-trivial solutions are a
more delicate issue.
\end{enumerate}
\begin{figure}
\centering
\subfloat[Function $\Psi^\mu(2K\cdot)$ for $\mu=\delta_0$, $K=2$.]{\includegraphics[width=6cm]{phi0p1.eps}
\label{subfig:fixedpointPsi mu delta0}}
\quad\subfloat[Function $\Psi^\mu(2K\cdot)$ for $\mu=\frac12\left( \delta_{-1} + \delta_1\right)$,
$K=3.5$.]{\includegraphics[width=6cm]{phimu.eps}
\label{subfig:fixedpointPsi mu}}
\caption{Plot of the fixed-point function $\Psi^\mu(2K\cdot)$ for two choices of $K$ and $\mu$.
$\Psi^{\delta_{0}}(\cdot) $ is strictly concave with derivative at the origin equal to $1/2$ (Fig. \ref{subfig:fixedpointPsi mu delta0}) but
even for a simple instance of $\mu$
(Fig. \ref{subfig:fixedpointPsi mu}) concavity is lost and there are
several non-trivial fixed-points, each of them corresponding to one circle of non trivial stationary solutions.
Note that in the case of Fig. \ref{subfig:fixedpointPsi mu}, $K < \widetilde{K}=5$ so that the phase transition is not given by
the derivative of $\Psi^\mu(2K\cdot)$ at the origin.}%
\label{fig:fixed point Psi}%
\end{figure}
\subsection{An overview of the results we present}
Here are two natural questions:
\begin{itemize}
\item What are the stability properties of the non-trivial stationary solutions?
\item What happens if $\mu$ is not symmetric?
\end{itemize}
Our work addresses these two questions and provides complete answers for weak disorder.
The precise set-up of our work is better understood if we remark from now that we can
assume $m_\omega:=\int \omega \mu (\dd \omega)=0$. In fact, if this is not the case we can map the model
to a model with $m_\omega=0$ by putting ourselves on the frame that rotates with speed $m_\omega$,
that is if we consider the diffusion $\{\varphi^\omega _{j}(t)- m_\omega t\}_{j=1, \ldots, N}$. So, we assume henceforth $m_\omega=0$
and we rewrite the natural frequencies as $\delta \omega$, with $\delta$ a non-negative parameter. We assume moreover that
\begin{equation}
\label{eq:suppHyp}
\Supp(\mu)\subseteq [-1, 1]\, .
\end{equation}
In this set-up, \eqref{FKP kuramoto disorder} becomes
\begin{equation}
\label{FKP kuramoto disorder delta}
\partial_t p_t^\delta(\theta,\omega)\, =\, \frac{1}{2} \Delta p_t^\delta(\theta,\omega) -\partial_\theta \Big(p_t^\delta(\theta,\omega)(\langle
J*p_t^\delta\rangle_\mu(\theta) +\delta\omega)\Big)\, .
\end{equation}
Note that this leads to (obvious) changes to \eqref{eq:qhatom}-\eqref{eq:fixedpointom}.
We have introduced this parameterization because the results that we present are for
small values of $\delta$.
In particular we are going to show that for any $K>1$, there exists $\delta_0>0$ such that for $\delta \in [0, \delta_0]$
\begin{itemize}
\item there exists a solution $p_t^\delta(\theta,\omega)$
to \eqref{FKP kuramoto disorder delta} of the form $q( \theta - c_\mu (\delta) t)$, we show that $c_\mu (\delta) =O(\delta^3)$ and we
actually give an expression
for
$\lim_{\delta \searrow 0}c_\mu (\delta) / \delta^3$: this is a rotating wave (or limit-cycle) for the dynamical system \eqref{FKP kuramoto disorder delta}
and we establish its stability under perturbations;
\item when $\mu$ is symmetric and $K> \widetilde K$ we show that there is, up to rotation symmetry, only one non-trivial solution and that it is (linearly and non-linearly) stable.
\end{itemize}
\medskip
The results we obtain are based on the rather good understanding that we have of the case $\delta=0$ that,
as we have already explained, is reversible and the corresponding Fokker-Planck PDE is of gradient flow type (e.g. \cite{cf:Otto} and references therein).
These properties have been exploited in \cite{cf:BGP} in order to extract a number of properties of
the Fokker-Planck PDE (denoted from now on: reversible PDE)
\begin{equation}
\label{eq:revK}
\partial_t p_t(\theta)\, =\, \frac{1}{2} \Delta p_t(\theta) -\partial_\theta \Big(p_t(\theta)( J*p_t)(\theta) \Big) \, ,
\end{equation}
and notably the linear stability of the non-trivial stationary solutions. In fact one can find in \cite{cf:BGP}
an analysis of the evolution operator linearized around the non-trivial stationary solutions. Some of the results
in \cite{cf:BGP} are recalled in the next section, but they are not directly applicable because
the $\delta=0$ case that corresponds to what interests us is rather
\begin{equation}
\label{FKP kuramoto disorder without drift}
\partial_t p_t(\theta, \omega)\, =\, \frac{1}{2} \Delta p_t(\theta, \omega) -\partial_\theta \Big(p_t(\theta, \omega)(\langle J*p_t\rangle_\mu(\theta) \Big) \, ,
\end{equation}
which we call {\sl non-disordered PDE}.
So the {\sl natural frequencies} have no effective role
beyond separating the various rotators into populations with given natural (ineffective) frequency
that now are just labels.
But in order to set-up a proper
perturbation procedure we need to control \eqref{FKP kuramoto disorder without drift} and, in particular,
we need (and establish) a spectral gap inequality for the evolution \eqref{FKP kuramoto disorder without drift} linearized around the non-trivial solutions.
This spectral analysis is going to be central both for the general and for the symmetric disorder case. In the general set-up we are going to exploit the
{\sl normally hyperbolic structure} \cite{cf:HPS,cf:SellYou} of the manifold of stationary solutions of \eqref{FKP kuramoto disorder without drift}
and the robustness of such structures (like in \cite{cf:GPPP}). In the case of
symmetric $\mu$ we can get more precise results by
ad hoc estimates, made possible by the explicit expressions
\eqref{eq:qhatom}-\eqref{eq:fixedpointom}, and use results in the general theory of operators \cite{Pazy1983} and
perturbation theory of self-adjoint operators \cite{Kato1995}.
The normal hyperbolic manifold approach allows to treat cases that are substantially more general and notably the case
of
\begin{equation}
\label{eq:AR}
\partial_t p_t(\theta,\omega)\, =\, \frac{1}{2} \Delta p_t(\theta,\omega) -\partial_\theta \Big(p_t(\theta,\omega)(\langle
J*p_t\rangle_\mu(\theta) +\delta U(\theta, \omega))\Big)\, ,
\end{equation}
which is the large $N$ limit of \eqref{eq:Kmod} with the term $\omega_j \dd t$ replaced by
$U(\varphi^\omega_j (t), \omega_j) \dd t$, with $U \in C^1( {\ensuremath{\mathbb S}} \times {\ensuremath{\mathbb R}} ; \, {\ensuremath{\mathbb R}} )$.
In this case each oscillator has its own non-trivial dynamics which may be very different from
the dynamics of other oscillators: consider for example
\begin{equation}
U(\varphi, \omega)\, =\,b+ \omega + a\sin (\varphi)\, , \ \ \ a, b \in {\ensuremath{\mathbb R}} \, ,
\end{equation}
and $\mu$ uniform over $[-1,1]$. For $a \in (-1,1)$ there are some {\sl active rotators}
\cite{cf:shinomoto1986a,cf:GPPP}
that in absence of noise and interaction ($\sigma=K=0$) rotate (this happens if $\vert b+\omega\vert >\vert a\vert$ and of course
the direction of rotation depends on the sign of $b+\omega$)
and others that instead are stuck at a fixed point (this happens if $\vert b+\omega\vert \le \vert a\vert$).
Our approach allows us to establish that there is a synchronization regime
for $K>1$ and $\delta$ small and to describe the dynamics of the system in this regime.
This is going to be detailed
in Section~\ref{sec:AR}.
\medskip
The two questions raised at the beginning of this section have been already repeatedly approached but looking at synchronized solutions as
bifurcation from incoherence. The results are hence for $K$ close to the critical value corresponding to the breakdown of linear stability of $1/2\pi$:
one can find a detailed review of the vast literature on this issue in \cite[Sec. III]{cf:acebron}.
Our results are instead for arbitrary $K>1$, but $\delta$ smaller than $\delta_0(K)$ and of course $\delta_0(K)$ vanishes as $K$ approaches $1$.
\section{Mathematical set-up and main results}
\label{sec:mainresults}
\subsection{The reversible and the non-disordered PDE}
\label{subsec:organizednondisorderedcase}
We first recall some results about the reversible PDE \eqref{eq:revK}.
The stationary solutions $q_0(\theta)=q(\theta,0)$ are, up to rotation invariance, given by
\eqref{eq:qhatom}-\eqref{eq:fixedpointom}, but formulas get simpler, namely
\begin{equation}
\label{eq:defstationarysolution nodisorder}
q_0(\theta)\, =\, \frac 1{Z_0(2Kr_0)} \exp(2Kr_0 \cos(\theta))\, ,
\end{equation}
where $Z_0(x):= Z(0, x)^\frac 12$ and
this time we have the more explicit expression $Z_0(x)\, =\, \int_{\bbS} e^{x\cos(\theta)} \dd\theta = 2\pi I_0(x)$ is the normalization
constant and $r_0$ is a solution of
the fixed-point problem
\begin{equation}
\label{eq:deffixedpoint nodisorder}
r_0\, =\, \Psi_0 (2Kr_0) \qquad \text{where} \qquad \Psi_0(x)\, :=\, \frac{I_1(x)}{I_0(x)}\, ,
\end{equation}
where we used standard notations for the modified Bessel functions
\begin{equation}
\label{def bessel}
I_i(x)\, =\, \frac 1{2\pi}\int_{{\ensuremath{\mathbb S}} } (\cos(\theta))^i\exp(x\cos(\theta))\dd\theta\, \qquad i=0,1\, .
\end{equation}
The mapping $\Psi_0$ is increasing, concave (see \cite{cf:Pearce}) and with derivative at $0$ equal to $\frac 12$. Consequently if
$K\leq
1$, $r_0=0$ is
the unique solution of the fixed-point problem, and $q(\cdot)\equiv\frac{1}{2\pi}$ is the only stationary solution of \eqref{eq:revK}.
If $K>1$, we get in addition a circle (because of the rotation invariance) of nontrivial stationary solutions
\begin{equation}
M_{\text{rev}}\, :=\, \{q_{\psi, 0}(\cdot):=q_0(\cdot-\psi):\, \psi\in{\ensuremath{\mathbb S}} \} \qquad \text{with} \qquad q_0(\theta)\,
:=\,\frac{\exp(2Kr_0\cos(\theta))}{\int_{{\ensuremath{\mathbb S}} }\exp(2Kr_0\cos(\theta))}
\end{equation}
where $r_0=r_0(K)$ is the unique non trivial fixed-point \eqref{eq:deffixedpoint nodisorder}.
\medskip
Let us now focus on the non-disordered PDE \eqref{FKP kuramoto disorder without drift}
and let us insist on the fact that we are interested in solutions such
that $\varphi^\delta_t(\cdot, \omega)$ is a probability density.
Observe then that if $q(\theta,\omega)$ is a stationary solution of \eqref{FKP kuramoto disorder without drift}, we see (Appendix
\ref{sec:appendix regularity pt with disorder}) that $q$ is
$C^\infty$ with respect to $\theta$ and that $\langle q \rangle_\mu$ is a stationary solution for \eqref{eq:revK}. So there
exists $\psi\in{\ensuremath{\mathbb S}} $ such that $\langle q\rangle_\mu = q_\psi$ and a short computation leads to
\begin{equation}
\langle J*q \rangle_\mu(\theta)\, =\, -K\sin(\theta-\psi)\, ,
\end{equation}
and, since
$\int_{\bbS} q(\theta,\omega) \dd \theta =1$ for almost all $\omega$, we obtain
that $q(\cdot,\omega)=q_{ \psi}(\cdot)$ for almost all $\omega$. In conclusion, with some abuse of notation, we can say the stationary solutions of \eqref{eq:revK} and
\eqref{FKP kuramoto disorder without drift} are the same: of course in the second case
the function space includes the dependence on $\omega$, so
we choose a different notation, that is $M_0$, for the corresponding
circle of non-trivial stationary solutions.
\medskip
An important issue for us is the stability of $M_0$ (for its existence we are assuming $K>1$) and for this
we denote by $A$ the linearized evolution operator of
\eqref{FKP kuramoto disorder without drift} around $q_0$
\begin{equation}
\label{def A}
Au(\theta,\omega)\, :=\, \frac 12 \Delta u(\theta,\omega) - \partial_\theta \Big(q_0(\theta) \langle J*u \rangle_\mu(\theta) + u(\theta,\omega)
J*q_0(\theta)\Big)
\end{equation}
with domain
\begin{equation}
\label{eq:defdomain A}
{\ensuremath{\mathcal D}} (A)\, :=\, \left\{u\in C^2({\ensuremath{\mathbb S}} \times{\ensuremath{\mathbb R}} ,{\ensuremath{\mathbb R}} ):\, \int_{\bbS} u(\theta,\omega)\dd\theta=0\, \text{ for all }\omega\right\}\, .
\end{equation}
For any smooth positive function $k:{\ensuremath{\mathbb S}} \mapsto {\ensuremath{\mathbb R}} $, we introduce the Hilbert space $H^{-1}_{k,\mu}$ defined by the closure of
${\ensuremath{\mathcal D}} (A)$ for the norm $\Vert \cdot \Vert_{-1,k,\mu}$ associated with the scalar product
\begin{equation}
\label{def scalar product sobolev with wage and mu}
\langle u,\, v \,\rangle_{-1,k,\mu}\, :=\, \int_{\bbR\times{\bbS }} \frac{{\ensuremath{\mathcal U}} (\theta,\omega){\ensuremath{\mathcal V}} (\theta,\omega)}{k(\theta)}\dd\theta \mu(\dd\omega),
\end{equation}
where $\omega$ a.s., ${\ensuremath{\mathcal U}} (\cdot,\omega)$ is the primitive of $u(\cdot,\omega)$ such that
$\int_{{\ensuremath{\mathbb S}} }\frac{{\ensuremath{\mathcal U}} (\theta,\omega)}{k(\theta)}\dd\theta =0$, and ${\ensuremath{\mathcal V}} (\cdot,\omega)$ is defined in the analogous fashion. Let us remark (see
\cite[Sec.~2]{cf:GPPP})
immediately that
\begin{equation}
\Vert u \Vert_{-1,k_1,\mu}^2\, \leq \, \frac{\Vert k_2 \Vert_\infty}{ \Vert k_1 \Vert_\infty}\Vert u \Vert_{-1,k_2,\mu}^2\, ,
\end{equation}
so that all the norms we have introduced are equivalent. For the case $k(\cdot) \equiv
1$ we use the notations $H^{-1}_\mu$ and
$\Vert\cdot\Vert_{-1,\mu}$.
We will prove the following result, which is just technical, but it will be of help
to understand our main results:
\medskip
\begin{proposition}
\label{th:spectral gap A}
A is essentially self-adjoint in $H^{-1}_{{q_0},\mu}$. Moreover the spectrum lies in $(-\infty,0]$, $0$ is a
simple eigenvalue, with eigenspace spanned by $\partial_\theta q_0$,
and there is a spectral gap, that is
the distance $\gl_{K}$ between $0$ and the rest of the spectrum is positive.
\end{proposition}
\medskip
The proof of this result builds on \cite[Th.~1.8]{cf:BGP} that deals with the
reversible case and the (lower) bound
on the spectral gap $\gl_{K}$ that we obtain coincides with the quantity $\lambda(K)$
in \cite[Th.~1.8]{cf:BGP} (this bound can be improved
as explained in in \cite[Sec.~2.5]{cf:BGP} and sharp estimates on the spectral gap
can be obtained in the limit $K \searrow 1$ and $K \nearrow \infty$).
For the reversible evolution, the
linear operator $L_{q_0}$ is defined by
\begin{equation}
\label{def Lq0}
L_{q_0}u(\theta)\, :=\, \frac 12 \Delta u(\theta) -\partial_\theta\Big( q_0(\theta)J*u(\theta)+u(\theta)J*q_0(\theta)\Big),
\end{equation}
with domain $D(L_{q_0})$ given by the $C^2({\ensuremath{\mathbb S}} ,{\ensuremath{\mathbb R}} )$ functions with zero integral.
\subsection{Synchronization: the main result without symmetry assumption}
Proposition~\ref{th:spectral gap A} is a key ingredient for our main results and
the functional space $H^{-1}$ appears in it, but an important role
is played also by $L^2(\lambda \otimes \mu)$, $\lambda$ is the Haar measure on ${\ensuremath{\mathbb S}} $,
whose norm is denoted by $\Vert \, \cdot \, \Vert_{2, \mu}$. For $C>0$ and $M \subset L^2(\lambda \otimes \mu)$ we set
${\ensuremath{\mathcal N}} _{2, \mu}(M, C):= \{u: \,$there exists $v\in M$ such that $\Vert u-v \Vert_{2, \mu} \le C\}$.
In the statement below $q\in M_0$ is the element of the manifold such that $q(\cdot,\omega)=
q_0(\cdot)$, cf. \eqref{eq:defstationarysolution nodisorder}, with $r_0(K)>0$ (hence $K>1$).
\medskip
\begin{theorem}
\label{th:expansion speed}
For every $K>1$
there exists $\delta_0=\delta_0(K)>0$ such that for $\vert \delta \vert \le \delta_0$
there exists $\widetilde q _\delta \in L^2(\lambda \otimes \mu)$, satisfying
$\Vert \widetilde q _\delta -q \Vert_{2, \mu}=O(\delta)$ and a value
$c_\mu (\delta) \in {\ensuremath{\mathbb R}} $ such that if we set
\begin{equation}
\label{eq:main1.1}
q^{(\psi)}_t (\theta, \omega)\, :=\, \widetilde q _\delta(\theta -c_\mu (\delta) t-\psi)\, ,
\end{equation}
then $q^{(0)}_t$ solves \eqref{FKP kuramoto disorder delta}. Moreover
\begin{enumerate}
\item
the family of solutions $\{q^{(\psi)}_\cdot\}_\psi$
is stable in the sense that there exist two positive constants $\beta=\beta(K)$ and $C=C(K)$
such that if $p_0^\delta \in {\ensuremath{\mathcal N}} _{2, \mu}(M_0,\delta)$, and $\int_{\ensuremath{\mathbb S}} p_0^\delta
(\theta, \omega) \dd \theta=0$ $\mu(\dd \omega)$-a.s., then there exists
$\psi_0\in {\ensuremath{\mathbb S}} $ such that for all $t\ge 0$
\begin{equation}
\Vert q_t^{(\psi_0)}-p_t^\delta \Vert_{2, \mu}\, \le \, 2 C\exp(-\beta t)\, .
\end{equation}
\item
we have
\begin{equation}
\label{eq:main1.2}
c_\mu(\delta)\, =\,
\delta^3
\frac{
\left\langle \omega\partial_\theta n^{(2)} , \partial_\theta q_0 \right \rangle_{-1, {q_0},\mu} }
{ \left\langle \partial_\theta q_0 ,\partial_\theta q_0
\right \rangle_{-1, {q_0}} } + O(\delta^5)\, ,
\end{equation}
where $n^{(2)}$ is the unique solution of
\begin{equation}
An^{(2)}\, =\,
\omega
\partial_\theta n^{(1)} \quad\text{ and } \quad \left\langle n^{(2)}, \partial_\theta q_0 \right \rangle_{-1, {q_0},\mu}
\, =\, 0\, ,
\end{equation}
and $n^{(1)}$ is the unique solution of
\begin{equation}
A n^{(1)}\, =\,
\omega \, \partial_\theta q_0 \quad\text{ and } \quad \crosqmu {n^{(1)}}{ \partial_\theta q_0 }\, =\, 0\, .
\end{equation}
\end{enumerate}
\end{theorem}
\medskip
In the proof of Theorem~\ref{th:expansion speed} one finds also further estimates, in
particular (see \eqref{eq:n_2}) that one has
\begin{equation}
\label{eq:qgddevel}
\tilde q_\delta \, =\, q_0 + \delta n^{(1)} + \delta^2 n^{(2)} + O_{L^2}(\delta^3)\, .
\end{equation}
Actually, see Remark~\ref{rem:pa}, the argument of proof can be pushed farther
to obtain arbitrarily many terms in development \eqref{eq:qgddevel}, as well as in
\begin{equation}
c_\mu (\delta)\, =\, c_3 \delta^3+ c_5 \delta^5 +\ldots \, .
\end{equation}
In Table~\ref{tab:1} we report a comparison between
the $c_\mu(\delta)$ obtained by solving numerically \eqref{FKP kuramoto disorder delta}
and by evaluating the leading order $c_3$, i.e. by using \eqref{eq:main1.2}.
\begin{table}[h!]
\begin{center}
\begin{tabular}{ | c | c | c | c | }
\hline \phantom{m}
$\delta$ & $K=2$ & $ K=1.5 $ & $ K=1.1 $ \\ \hline
$0.5$ & $-1.56300 \cdot 10^{-2}$ & $-8.59626 \cdot 10^{-2}$ & $-3.01064 \cdot 10^{-1}$ \\ \hline
$0.1$ & $-1.23998 \cdot 10^{-2}$ & $-6.84835 \cdot 10^{-2}$ & $-2.72117 \cdot 10^{-1}$ \\ \hline
$0.05$ & $-1.23072 \cdot 10^{-2}$ & $-6.79553 \cdot 10^{-2}$ & $-2.69460 \cdot 10^{-1}$ \\ \hline
$0.01$ & $-1.22776 \cdot 10^{-2}$ & $-6.77921 \cdot 10^{-2}$ & $-2.68603 \cdot 10^{-1}$ \\ \hline
$0.005$& $-1.22767 \cdot 10^{-2}$ & $-6.77869 \cdot 10^{-2}$ & $-2.68576 \cdot 10^{-1}$ \\ \hline
\multicolumn{4}{c}{} \\ \hline
$c_3$ & $-1.22764 \cdot 10^{-2}$ & $-6.77851 \cdot 10^{-2}$ & $-2.68567 \cdot 10^{-1}$ \\ \hline
\end{tabular}
\caption{\label{tab:1} For the case
$
\mu= p \delta_{1-p} + (1-p) \delta_{-p}$, $p=0.2$, we have computed (numerically)
$c_\mu(\delta)/\delta^3$ for three values of $K$
and five values of $\delta$. In the last line we report the value
$c_3= \lim_{\delta\searrow 0} c_\mu(\delta)/\delta^3$ that one obtains by using \eqref{eq:main1.2}. }
\end{center}
\end{table}
\subsection{Symmetric disorder case}
Let us focus on the case in which the
distribution of the disorder $\mu$ is symmetric.
In this case, at least for small disorder,
Theorem~\ref{th:expansion speed} is just telling us that the leading order
in the development for the speed $c_\mu(\delta)$ is zero: one can actually work harder and show
that such a development yields zero terms to all orders.
In reality in this case we already know, see \eqref{eq:qhatom}-\eqref{eq:fixedpointom},
that for $K$ sufficiently large there is at least a non-trivial stationary profile, hence, by
rotation symmetry, at least one whole circle of stationary solutions.
Actually, we can show that for $\delta$ small there is just one circle,
that we call $M_\delta$, of non-trivial stationary solutions
and this circle converges to $M_0$ as $\delta \searrow 0$ (in $C^j$, for every $j$) so
the rotating solutions found in Theorem~\ref{th:expansion speed} must be the stationary solutions
in $M_\delta$.
In order to be precise about this issue, we point out that
\eqref{eq:qhatom}-\eqref{eq:fixedpointom} are written for
\eqref{FKP kuramoto disorder}
while we work rather with
\eqref{FKP kuramoto disorder delta}. The changes are obvious, but we introduce a notation
for the analog of \eqref{eq:fixedpointom}:
\begin{equation}
\label{eq:deffixedpoint disorder}
r_\delta\,=\, \Psi_\delta^\mu(2Kr_\delta), \qquad \text{where,}\, \Psi_\delta^\mu(x)\,:=\, \int_{\bbR}\frac{\int_{\bbS}\cos(\theta)S(\theta, \delta\omega,
x)\dd\theta}{Z(\delta\omega, x)}\mu(\dd\omega)\, .
\end{equation}
\medskip
\begin{lemma}
\label{th: Psimu concave}
For all $K_{\min}<K_{\max}$, there exists $\delta_1=\delta_1(K_{\min}, K_{\max})>0$ such that, for all $0<K_{\min}<K<K_{\max}$
and all $\delta\leq\delta_1$ the function $\Psi_\delta^\mu$ is strictly concave on $[0,1]$. Therefore for
\eqref{eq:fixedpointom} has only a positive solution $r_\delta=r_\delta(K, \mu)$. Moreover $
\lim_{\delta \searrow 0} r_\delta= r_0$.
\end{lemma}
\medskip
\medskip
We point out that in spite of the fact that $\Psi^\mu$ is explicit (cf. \eqref{eq:deffixedpoint nodisorder}), it is not so straightforward to
show that it is concave. We show that $\Psi_\delta^\mu$ remains strictly concave for a small $\delta$ via a perturbation argument. But the conjecture
(see \cite{cf:dH} and \cite{cf:dPdH}) that $\Psi^\mu$ is strictly concave for unimodal distributions $\mu$ is still an open issue.
\begin{rem}
\rm
A direct computation shows that the derivative of $\Psi^\mu_\delta$ at the origin is $1/(2\widetilde K _\delta)$, for
$\widetilde K _\delta\,:=\,
\left(\int_{\bbR} \frac{\mu(\dd\omega)}{1+4\delta^2\omega^2}\right)^{-1}$
(of course $\widetilde K_1$ coincides with $\widetilde K$,
introduced in \eqref{eq:Ktilde}).
Under the hypothesis of Lemma \ref{th: Psimu concave}, one therefore
sees that there is a synchronization transition at $K=\widetilde K _\delta$ in the sense
that for $K\le \widetilde K _\delta$ the only stationary solution is $\frac 1{2\pi}$
while for $K>\widetilde K _\delta$ also the manifold of non-trivial stationary solutions
appears (and there is no other stationary solution).
\end{rem}
\medskip
Theorem~\ref{th:expansion speed} provides a
stability statement for $M_\delta$. This result can be sharpened and for this let us introduce
the linear operator
\begin{equation}
\label{eq:defLqmu}
L^\omega _{{q}}u(\theta, \omega)\, :=\, \frac 12 \Delta u(\theta, \omega) - \partial_{\theta} \left(
u(\theta, \omega)\left(
\langle J \ast {q}\rangle_\mu(\theta) + \delta\omega\right) + {q}(\theta, \delta\omega) \langle J\ast u\rangle_\mu(\theta)
\right),
\end{equation}
The domain ${\ensuremath{\mathcal D}} (L^\omega_{{q}})$ of the operator $L^\omega_{{q}}$ is chosen to be the same as for $A$, cf. \eqref{eq:defdomain A}.
\medskip
We place ourselves within the framework of Lemma \ref{th: Psimu concave}, in the sense that $\delta$ is small enough to ensure
the uniqueness of a non-trivial stationary solution
(of course existence requires $K> \tilde K_\delta$ and this is implied
by $K>1$ if $\delta$ is sufficiently small).
We prove a number of properties of the linear operator \eqref{eq:defLqmu}, saying notably that
it has a simple eigenvalue at zero and
the rest of spectrum is at a positive distance from zero and it is in a cone in that lies in the negative complex half plane. We summarize in the next statement the qualitative features
of our results on $L_q^\omega$, but what we really prove are quantitative explicit estimates:
the interested reader finds them in Section~\ref{sec:sym}.
\medskip
\begin{theorem}
\label{th:spectral prop L disorder}
The operator $L^\omega_{{q}}$ has the following spectral properties:
$0$ is a simple eigenvalue for $L^\omega_{{q}}$, with eigenspace spanned by $(\theta, \omega)\mapsto q'(\theta, \omega)$.
Moreover, for all $K>1$, $\rho\in(0,1)$, $\alpha\in(0,\pi/2)$, there exists $\delta_2=\delta_2(K, \rho, \alpha)$ such that for all $0\leq
\delta\leq \delta_2$, the
following is true:
\begin{itemize}
\item $L^\omega_{{q}}$ is closable and its closure has the same domain as the domain of the self-adjoint extension of $A$;
\item The spectrum of $L^\omega_{{q}}$ lies in a cone $C_{\alpha}$ with vertex $0$ and angle $\alpha$
\begin{equation}
C_{\alpha}\, :=\, \ens{\lambda\in{\ensuremath{\mathbb C}} }{\frac{\pi}{2} + \alpha\leq
\arg(\lambda)\leq \frac{3\pi}{2}-\alpha}\subseteq \ens{z\in{\ensuremath{\mathbf C}} }{\Re(z)\leq 0}\, ;
\end{equation}
\item There exists $\alpha'\in(0, \frac\pi2)$ such that $L^\omega_{{q}}$ is the infinitesimal generator of an analytic semi-group defined on a
sector $\{\lambda\in{\ensuremath{\mathbb C}} ,\, |\arg(\lambda)|< \alpha'\}$;
\item The distance between $0$ and the rest of the spectrum is strictly positive and is at least equal to
$\rho\gl_{K}$, where $\gl_{K}$ is the spectral gap of the operator $A$ introduced in Proposition \ref{th:spectral
gap A}.
\end{itemize}
\end{theorem}
\subsection{Organization of remainder of the paper}
In Section~\ref{sec:hyperbolic structure}
we introduce the notion of stable normally hyperbolic manifold,
we recall its robustness properties,
and show that $M_0$ is in this class of manifolds.
The essential ingredient is Proposition~\ref{th:spectral gap A}
that, directly or indirectly, plays a role in each subsequent section.
Section~\ref{sec:hyperbolic structure} is also devoted to
the proof of Proposition~\ref{th:spectral gap A}.
The proof of
Theorem~\ref{th:expansion speed} is then completed
in Section~\ref{sec:pa}, that is mainly devoted to perturbation
arguments. The case of the active rotators is treated in
Section~\ref{sec:AR}, while
Section~\ref{sec:sym} deals with the case symmetric disorder distribution and,
notably, with the proof of Theorem~\ref{th:spectral prop L disorder}
and of a number of related quantitative estimates.
\section{Hyperbolic structures and periodic solutions}
\label{sec:hyperbolic structure}
In this section we present the arguments proving the existence of the periodic solution of Theorem \ref{th:expansion speed}. We rely on the fact
that the circle of stationary solutions $M_0$ is a stable normally hyperbolic manifold, and on the robustness of this kind of structure : adding
the perturbation term $-\delta\partial_\theta(p_t(\theta,\omega)\omega)$ in \eqref{FKP kuramoto disorder without drift}, this manifold $M_0$ is deformed into
another manifold $M_\delta$, and thanks to the rotation invariance of the problem, $M_\delta$ is a circle too. The spectral gap of operator $A$
(Property \ref{th:spectral gap A}) which induces the hyperbolic property of $M_0$ is proved at the end of this section.
\subsection{Stable normally hyperbolic manifolds}
We start by quickly reviewing
the notion of of stable normally hyperbolic manifold (SNHM).
The evolution of \eqref{FKP kuramoto disorder delta} will be studied in the space $X^1_\mu$ defined by
\begin{equation}
X^1_\mu\, :=\, \left\{u\in L^2(\lambda\otimes\mu),\, \int_{{\ensuremath{\mathbb S}} }u(\theta,\omega)\dd\theta=1\quad \omega\text{ a.s.}\right\}
\end{equation}
where $\lambda$ denotes the Lebesgue measure on ${\ensuremath{\mathbb S}} $. This is made possible by the conservative character of the dynamics. The
$L^2$-norm with respect to the measure $\lambda\otimes\mu$ will be denoted by $\Vert \cdot\Vert_{2,\mu}$. We will also use the space
$X^0_\mu$ defined by
\begin{equation}
X^0_\mu\, :=\, \left\{u\in L^2(\lambda\otimes\mu),\, \int_{{\ensuremath{\mathbb S}} }u(\theta,\omega)\dd\theta=0\quad \omega\text{ a.s.}\right\}\, .
\end{equation}
To define a SNHM, we need a dynamics: we have in mind \eqref{FKP kuramoto disorder delta}
but for the moment let us just think of an evolution semigroup in $X^1_\mu$ that gives rise to $\{ u_t\}_{t\ge 0}$, with $u_0=u$,
to which we can associate a linear evolution semigroup $\{\Phi(u, t)\}_{t \ge 0}$ in $X^0_\mu$, satisfying
$\partial_t \Phi(u, t)v =L(t) \Phi(u, t)v$ and $\Phi(u, 0)v=v$, where $L(t)$ is the operator
obtained by linearizing the evolution around $u_t$.
For us a SNHM $M\subset X^1_\mu$ (in reality we are interested only
in $1$-dimensional manifolds, that is curves, but at this stage this does not really play a role)
of characteristics $\lambda_1$, $\lambda_2$ ($0\le \lambda_1< \lambda_2$) and $C>0$ is
a $C^1$ compact connected manifold which is invariant under the dynamics
and for every $u \in M$ there exists a projection $P^o(u)$ on the
tangent space of $M$ at $u$, that is ${\ensuremath{\mathcal R}} (P^o(u))=:T_uM$, which, for $v \in L^2_0$, satisfies the following properties:
\begin{enumerate}
\item for every $t\ge 0$ we have
\begin{equation}
\Phi(u,t) P^o (u_0)v\, =\, P^o (u_t)\Phi(u,t) v\, ,
\end{equation}
\item we have
\begin{equation}
\label{eq:shyp1}
\Vert \Phi(u,t) P^o(u_0) v\Vert_{2,\mu} \, \le \, C \exp( \lambda_1 t)
\Vert v \Vert_{2,\mu}\, ,
\end{equation}
and, for
$P^s\, :=\, 1-P^o$, we have
\begin{equation}
\label{eq:shyp2}
\Vert \Phi(u,t) P^s(u_0) v\Vert_{2,\mu} \, \le \, C \exp( -\lambda_2 t)
\Vert v \Vert_{2,\mu}\, ,
\end{equation}
for every $t\ge 0$;
\item
there exists a negative continuation of
the dynamics $\{ u_t\}_{t \le 0}$ and of the linearized
semigroup
$\{ \Phi(u, t) P^o(u_0) v \}_{t \le 0}$ and for any such continuation
we have
\begin{equation}
\label{eq:shyp3}
\Vert \Phi(u,t) P^o(u_0) v\Vert_{2,\mu} \, \le \, C \exp( -\lambda_1 t)
\Vert v \Vert_{2,\mu}\, ,
\end{equation}
for $t \le 0$.
\end{enumerate}
\medskip
\subsection{$M_0$ is a SNHM}
First of all: the dynamics on $M_0$ is trivial. For $q_\psi\in M_0$, the projection $P^o_{q_\psi}$ on the tangent space is the
projection on the subspace spanned by $q^\prime_\psi$:
\begin{equation}
\label{def P^o}
P^o_{q_\psi} u\, =\, \frac{\left\langle u,q^\prime_\psi\right\rangle_{-1,q_\psi,\mu}}{\left\langle q^\prime_\psi,q^\prime_\psi\right\rangle_{-1,q_\psi}}q^\prime_\psi
\end{equation}
and since the dynamic on the manifold is trivial, we are allowed to choose for the parameters $\lambda_1=0$ and $\lambda_2=\gl_{K}$
(where we recall that $\gl_{K}$ is given by Proposition \ref{th:spectral gap A}).
\medskip
We are in the same situation as in \cite{cf:GPPP}. For a suitable perturbation and if $\delta$ is small enough, the circle $M_0$ is
smoothly
transformed into another SNHM $M_\delta$, which is close to $M_0$. The proof is the same as in \cite[Sec.~5]{cf:GPPP}, which, in turn builds
on results in \cite{cf:SellYou}): the spaces we are working in are more general since we have to deal with the disorder. Here suitable
perturbation means being an element of $C^1(X^0_\mu,H^{-1}_{\mu})$, but it is clearly the case for the perturbation $u\mapsto
-\delta\, \omega\, \partial_\theta u$ when $\mu$ is of compact support. The following theorem works for all $C^1(X^0_\mu,H^{-1}_{\mu})$
perturbations:
\medskip
\begin{theorem}
\label{th:M}
\cite[Sec.~5]{cf:GPPP}
For every $K>1$ there exists $\delta_0>0$
such that if $\delta \in [0, \delta_0]$ there exists a stable normally hyperbolic
manifold $M_\delta$ in $X^1_\mu$ for the perturbed equation \eqref{FKP kuramoto disorder delta}. Moreover
we can write
\begin{equation}
\label{eq:M}
M_\delta \, =\, \left\{q_\psi + \phi_\delta \left(q_\psi\right):\, \psi \in {\ensuremath{\mathbb S}} \right\}\, ,
\end{equation}
for a suitable function $\phi_\delta\in C^1(M_0,X^0_\mu)$ with the properties that
\begin{itemize}
\item $\phi_\delta (q) \in {\ensuremath{\mathcal R}} (A)$;
\item
there exists
$C>0$ such that
$\sup_{\psi }(\Vert \phi _\delta \left(q_\psi\right)\Vert_{2,\mu}+
\Vert \partial_\psi\phi_\delta(q_\psi)\Vert _{2,\mu}) \le C \delta$.
\end{itemize}
\end{theorem}
\medskip
\begin{rem}\label{rem:un}
\rm
A byproduct of the proof in \cite[Sec.~5]{cf:GPPP} is also that
$M_\delta$ is the unique invariant manifold in a $L^2(\lambda,\mu)$-neighborhood of $M_0$. So in the case of
\eqref{FKP kuramoto disorder delta}, thanks to the symmetry of the problem
that tells us that any rotation of $M_\delta$ is still a invariant manifold, $M_\delta$ is in fact a circle, and that the
dynamics on this circle is a traveling wave of constant (possibly zero) speed $c_\mu(\delta)$. So the invariant manifold we get for \eqref{FKP kuramoto disorder delta} is even $C^\infty$. In
this sense, when dealing with \eqref{FKP kuramoto disorder delta}, we are using
only part of the strength of Theorem~\ref{th:M}. Of course this symmetry argument does not
apply when dealing with \eqref{eq:AR}.
\end{rem}
\medskip
\begin{rem}\label{rem:stability}
\rm
Theorem \ref{th:M} addresses the existence and the linear stability of the manifold $M_\delta$.
The non-linear stability statement in Theorem~\ref{th:expansion speed}(1) follows from
Theorem~\ref{th:M}
combined with \cite[Theorem 8.1.1]{cf:Henry},
when the dynamics is periodic with non zero speed on $M_\delta$. If $M_\delta$ is a manifold of stationary points, the argument for the non-linear stability follows by repeating the argument
in
\cite[Th. 4.8]{cf:GPP}, where the non-disordered case is treated.
\end{rem}
\medskip
We now prove Proposition~\ref{th:spectral gap A} and thus that $M_0$ is a SNHM.
\medskip
\subsection{The spectral gap estimate (proof of Proposition \ref{th:spectral gap A})}
We start by remarking that $A$ is symmetric
for the scalar product $\crosqmu{\cdot}{\cdot}$ (recall \eqref{def scalar product sobolev with wage and mu}). In fact, for $u$ and
$v$ in ${\ensuremath{\mathcal D}} (A)$, a short computation gives (in the following we use the notation $u^\prime(\theta,\omega)=\partial_\theta u(\theta,\omega)$)
\begin{align}
\crosqmu{v}{Au} & \, =\, \int_{\bbR\times{\bbS }}
\left[\frac{{\ensuremath{\mathcal V}} (\theta,\omega)}{q_0(\theta)}\left(\frac{ u^\prime(\theta,\omega)}{2}-u(\theta,\omega)J*q_0(\theta)-q_0(\theta)\langle J*u
\rangle_\mu(\theta) \right) \right]\dd \theta\dd\mu\\ \notag
& \, =\, -\frac 12 \int_{\bbR\times{\bbS }} \frac{u(\theta,\omega)v(\theta,\omega)}{q_0(\theta)}\dd \theta\dd\mu+\int_{{\ensuremath{\mathbb R}} }\int_{({\ensuremath{\mathbb S}} )^2}
v(\theta,\omega)\tilde J*u(\theta,\omega')\dd\theta \dd\mu \otimes\mu\, ,
\end{align}
where $\tilde J(\theta)=K\cos(\theta)$. We now first prove an inequality for $A$
that is stronger than the spectral gap inequality and then deduce that $A$
is (essentially) self-adjoint. We define
the two following scalar products, which were used for the non-disordered case in \cite{cf:BGP}:
\begin{equation}
\label{eq:defNorm-1 non-des}
\crosq{u}{v}\, :=\, \int_{\bbS} \frac{{\ensuremath{\mathcal U}} (\theta){\ensuremath{\mathcal V}} (\theta)}{q_{0}(\theta)}\dd\theta \, ,
\end{equation}
where ${\ensuremath{\mathcal U}} (\cdot)$ is the primitive of $u(\cdot)$ such that $\int_{\bbS} \frac{{\ensuremath{\mathcal U}} (\theta)}{q_{0}(\theta)}\dd\theta=0$ and
\begin{equation}
\label{eq:defNorm2q non-des}
\cro{u}{v}_{2,q_0}\, :=\, \int_{\bbS} \frac{u(\theta)v(\theta)}{q_0(\theta)}\dd\theta\, .
\end{equation}
We denote the closures of ${\ensuremath{\mathcal D}} (L_0)$ for these scalar products respectively by $H^{-1}_{{q_0}}$ and $L^2_{{q_0}}$. In the disordered
case, $L^2_{{q_0}}$ corresponds to the space $L^2_{{q_0},\mu}$, which we define by the closure of ${\ensuremath{\mathcal D}} (A)$ with respect to the
norm $\Vert \cdot\Vert_{2,{q_0},\mu}$ associated with the scalar product
\begin{equation}
\label{def scalar product L2 with wage and mu}
\croLqmu{u}{v}\, :=\, \int_{\bbR} \int_{\bbS} \frac{u(\theta,\omega)v(\theta,\omega)}{q_0(\theta)} \dd\theta \dd\mu \, .
\end{equation}
The two Dirichlet forms for the disordered and non-disordered case
are respectively
\begin{equation}
\label{dirichlet disorder}
{\ensuremath{\mathcal E}} _\mu(u) \, =\, -\crosqmu{Au}{u}\, ,
\end{equation}
and
\begin{equation}
\label{dirichlet non-disorder}
{\ensuremath{\mathcal E}} (u)\, =\, -\crosq{ L_{q_0}u}{u} \, .
\end{equation}
As in \cite{cf:BGP}, we first prove a spectral gap type inequality that involves the scalar product
$\cro{\cdot}{\cdot}_{2,q_0}$. For this we introduce the projections on the line spanned by $q^\prime_0$ in the spaces $L^2_{{q_0},\mu}$ and
$L^2_{{q_0}}$
\begin{equation}
P_{2,{q_0},\mu} u=\frac{\croLqmu{u}{q^\prime_0}}{\croLq{q^\prime_0}{q^\prime_0}}q^\prime_0 \qquad \text{for all}
\,u=u(\theta,\omega)\in L^2_{{q_0},\mu} \, ,
\end{equation}
and
\begin{equation}
P_{2,{q_0}}u\, =\, \frac{\croLq{u}{q^\prime_0}}{\croLq{q^\prime_0}{q^\prime_0}}q^\prime_0 \qquad \text{for all}\
\in
L^2_{{q_0}} \, .
\end{equation}
Remark that since $q^\prime_0$ does not depend on $\omega$,
\begin{equation}
\croLqmu{q^\prime_0}{q^\prime_0}\, =\, \croLq{q^\prime_0}{q^\prime_0} \quad
\text{ and } \quad \crosqmu{q^\prime_0}{q^\prime_0}\, =\, \crosq{q^\prime_0}{q^\prime_0}\, ,
\end{equation}
and that for all $u\in L^2_{{q_0},\mu}$
\begin{equation}
\label{link projections}
P_{2,{q_0},\mu}u\, =\, \langle P_{2,{q_0}}u \rangle_\mu\, =\, P_{2,{q_0}}\langle u\rangle_\mu\, .
\end{equation}
\begin{proposition}
\label{prop:min dirichlet des}
For all $u\in L^2_{{q_0},\mu}$ such that for almost every $\omega$, $\int_{\bbS} u(\cdot, \omega) =0$
\begin{equation}
{\ensuremath{\mathcal E}} _{\mu}(u)\, \geq \, c_K \croLqmu{u-P_{2,{q_0},\mu}u}{u-P_{2,{q_0},\mu}u}\, ,
\end{equation}
with
\begin{equation}
c_K\, =\, 1-K(1-r_0^2)\in(0,1/2)\, .
\end{equation}
\end{proposition}
The proof of this proposition relies on the corresponding result for the non-disordered case.:
\begin{proposition}
(see
\cite[Prop. 2.3]{cf:BGP})
\label{prop:min dirichlet non-des}
For all $u\in L^2_{{q_0}}$ such that for almost every $\omega$, $\int_{\bbS} u(\cdot, \omega) =0$
\begin{equation}
{\ensuremath{\mathcal E}} (v)\, \geq \, c_K\croLq{ u-P_{2,{q_0}}u}{u-P_{2,{q_0}}u}\, .
\end{equation}
\end{proposition}
\medskip
\noindent
\textit{Proof of Proposition \eqref{prop:min dirichlet des}.}
The first step of the proof is to make the Dirichlet form of the non-disordered case appear in the the disordered case one, that is
\begin{align}
{\ensuremath{\mathcal E}} _\mu(u) & \, =\, \langle {\ensuremath{\mathcal E}} (u) \rangle_\mu +\int_{{\ensuremath{\mathbb R}} }\int_{({\ensuremath{\mathbb S}} )^2} u(\theta,\omega)\tilde J*[u(\theta,\omega)-u(\theta,\omega')]\dd\theta
\dd\mu\otimes \mu\\ \label{link Dirichlet}
& \, =\, \langle {\ensuremath{\mathcal E}} (u) \rangle_\mu +\frac 12 \int_{{\ensuremath{\mathbb R}} }\int_{({\ensuremath{\mathbb S}} )^2} [u(\theta,\omega)-u(\theta,\omega')]\tilde
J*[u(\theta,\omega)-u(\theta,\omega')]\dd \theta \dd\mu\otimes \mu\, ,
\end{align}
and from Proposition \eqref{prop:min dirichlet non-des} we see that
\begin{equation}
\langle {\ensuremath{\mathcal E}} (u)\rangle_\mu \, \geq \, c_K \croLq{ u-P_{2,{q_0}}u}{u-P_{2,{q_0}}u}\, .
\end{equation}
Now remark that if we define
\begin{equation}
v\, =\, u-P_{2,{q_0},\mu}u\, ,
\end{equation}
using \eqref{link projections} we get
\begin{equation}
v-P_{2,{q_0}}v\, =\, u-P_{2,{q_0}}u\, ,
\end{equation}
and so
\begin{equation}
\langle {\ensuremath{\mathcal E}} (u)\rangle_\mu \, \geq \, c_K \croLqmu{v-P_{2,{q_0}}v}{v-P_{2,{q_0}}v}\, .
\end{equation}
We now introduce an orthogonal decomposition of the space $L^2_{{q_0}}$ which is well adapted to the convolution with $\tilde J$.
\begin{lemma}
(See \cite[Lemma 2.1]{cf:BGP}.)
\label{lem:decomposition L2q}
We have the following decomposition
\begin{equation}
L^2_{{q_0}}\, =\, F_0\oplus^\bot F_{1/2}\oplus^\bot F_{K-1/2}
\end{equation}
where
\begin{equation}
F_0\, :=\, \left\{\theta\mapsto a_0+\sum_{j\geq 2}a_j\cos(j\theta)+b_j\sin(j\theta)\, ;\, \sum_j a_j^2+b^2_j<\infty \right\}
\end{equation}
and both $F_{1/2}$ and $F_{K-1/2}$ are one dimensional subspaces generated respectively by $\theta\mapsto \sin(\theta)q(\theta)\,
(=-q^\prime_0(\theta)/2Kr_0)$ and by $\theta \mapsto \cos(\theta) q_0 (\theta)$. Moreover, when $u\in F_\lambda$, then
\begin{equation}
\tilde J *u\, =\, \frac{\lambda}{q_0} u\, .
\end{equation}
\end{lemma}
\medskip
With the help of Lemma~\ref{lem:decomposition L2q} we can find a lower bound for the last term in \eqref{link Dirichlet}: choose $\alpha$ such that $P_{2,{q_0}}u=\alpha q^\prime_0$, so that
we can write
\begin{equation}
\label{min Dirchlet}
{\ensuremath{\mathcal E}} _\mu(u) \, \geq \,c_K \croLqmu{ v-P_{2,{q_0}}v}{v-P_{2,{q_0}}v} + \frac{\croLq{q^\prime_0}{q^\prime_0}}{4} \int_{({\ensuremath{\mathbb S}} )^2}
(\alpha(\omega)-\alpha(\omega'))^2 \dd\mu\otimes \mu\, .
\end{equation}
But if $P_{2,{q_0}}v=\beta q^\prime_0$ (recall that $v=u-P_{2,{q_0},\mu}u$), then since $P_{2,{q_0},\mu}u$ is colinear to
$q^\prime_0$, for almost all $\omega$, $\omega'$
\begin{equation}
\beta(\omega)-\beta(\omega')\, =\, \alpha(\omega)-\alpha(\omega')
\end{equation}
and since $v$ is orthogonal to $q^\prime_0$ (with respect to $\langle \cdot,\cdot\rangle_{2,q_0,\mu}$) we get
\begin{equation}
\int_{{\ensuremath{\mathbb R}} } \beta(\omega)\dd\mu\, =\, 0 \, .
\end{equation}
So \eqref{min Dirchlet} becomes
\begin{equation}
{\ensuremath{\mathcal E}} _\mu(u) \, \geq \,c_K \croLqmu{v-P_{2,{q_0}}v}{v-P_{2,{q_0}}v} + \frac{\croLq{q^\prime_0}{q^\prime_0}}{2} \int_{{\ensuremath{\mathbb S}} } \beta^2(\omega) \dd\mu\,
.
\end{equation}
It is sufficient to compare this last minoration with the norm $\croLqmu{v}{v}$, and from Lemma \ref{lem:decomposition L2q} it comes
\begin{equation}
\croLqmu{v}{v}\, =\, \croLqmu{ v-P_{2,{q_0}}v}{v-P_{2,{q_0}}v} + \croLq{ q^\prime_0}{q^\prime_0} \int_{{\ensuremath{\mathbb S}} } \beta^2(\omega) \dd\mu\, .
\end{equation}
This completes the proof of Proposition~\ref{prop:min dirichlet des}.
\hfill $\quad \Box$ \bigskip
We now need two lemmas comparing the scalar products $\croLqmu{\cdot}{\cdot} $ and $\crosqmu{\cdot}{\cdot} $. They correspond to
Lemmas 2.4 and 2.5 in \cite{cf:BGP}. Their proofs are very similar to the proofs
of the results corresponding results in \cite{cf:BGP}
(to which we refer also for the explicit values of the constants $C$ and $c$ appearing below)
and they use in particular the rigged
Hilbert space representation of $H^{-1}_{{q_0},\mu}$ (see \cite[p.82]{MR697382}): namely, one can identify $H^{-1}_{{q_0},\mu}$ as
the dual space $V'$ of the space $V$ closure of ${\ensuremath{\mathcal D}} (A)$ with respect to the norm $\Vert{u}\Vert_V:= \left( \int_{{\ensuremath{\mathbb R}} \times{\ensuremath{\mathbb S}} } v'(\theta,
\omega)^2
\dd\theta\mu(\dd\omega) \right)^\frac12$. The pivot space $H$ is the usual
$L^2(\lambda\otimes\mu)$ (endowed with the Hilbert norm $\Vert u\Vert_{2,\mu}:=\left( \int_{\bbR\times{\bbS }} u(\theta,
\omega)^2\dd\theta\mu(\dd\omega) \right)^\frac12$). In particular, one easily sees that the inclusion $V\subseteq H$ is dense.
Consequently, one can define $T:H\rightarrow V'$ by setting $Tu(v)= \int_{\bbR\times{\bbS }} u(\theta, \omega) v(\theta, \omega)\dd\theta \mu(\dd\omega)$.
One can prove that $T$ continuously injects $H$ into $V'$ and that $T(H)$ is dense into $V'$ so that one can identify $u\in H$
with $Tu\in V'$. Then for $u\in H$,
\begin{equation}
\Vert u\Vert_{V'} = \Vert Tu\Vert_{V'} = \sup_{v\in V} \frac{\int {\ensuremath{\mathcal U}} v'}{\Vert v\Vert_{V}}= \sqrt{\int \frac{{\ensuremath{\mathcal U}} ^2}{q_0}}\, ,
\end{equation}
which enables us to identify $H^{-1}_{{q_0},\mu}$ with $V'$.
\medskip
We define the projection in
$H^{-1}_{{q_0},\mu}$:
\begin{equation}
P_{-1,{q_0},\mu}u\, =\, \frac{\crosqmu{u}{q^\prime_0}}{\crosq{q^\prime_0}{q^\prime_0}}q^\prime_0\, .
\end{equation}
\begin{lemma}
\label{lem:comparaison norm projections}
For every $K>1$ there exists a constant $C=C(K)>0$ such that
for $u\in L^2_\mu$ such that $\int_{\bbS} u =0$ for almost every $\omega$
\begin{equation}
\begin{split}
\croLqmu{u-P_{2,{q_0},\mu}u}{u-P_{2,{q_0},\mu}u}\, &\geq \, e^{4Kr_0}C
\croLqmu{u-P_{-1,{q_0},\mu}u}{u-P_{-1,{q_0},\mu}u}\\
&\geq \, C \crosqmu{u-P_{-1,{q_0},\mu}u}{u-p_{-1,{q_0},\mu}u}\, .
\end{split}
\end{equation}
\end{lemma}
\begin{lemma}
\label{lem:comparaison norm}
For every $K>1$ there exists $c=c(K)>0$ such that for $u\in L^2_\mu$ such that $\int_{\bbS} u =0$ for almost every $\omega$ and
\begin{equation}
\crosqmu{u}{u}\, \geq \, c \croLqmu{P_{2,{q_0},\mu}u}{P_{2,{q_0},\mu}u}\, .
\end{equation}
\end{lemma}
\medskip
\noindent
{\it Proof of Proposition~\ref{th:spectral gap A}.}
Of course Proposition \ref{prop:min dirichlet des} and Lemma \ref{lem:comparaison norm projections} imply directly the spectral
gap inequality for the Dirichlet form:
\begin{equation}
\label{low bound Dirichlet}
{\ensuremath{\mathcal E}} (u) \, \geq \, c_KC \crosqmu{u-P_{-1,{q_0},\mu}u}{u-P_{-1,{q_0},\mu}u}\qquad \text{for all}\, u\in H^{-1}_{{q_0},\mu}\, .
\end{equation}
We now prove the self-adjoint property of $A$. It is sufficient to prove that the range of $1-A$ is dense in $H^{-1}_{\mu}$ (see
\cite[p.113]{MR697382}).
For $u, v \in D(A)$, we have
\begin{multline}
\label{expr scal prod 1-A}
\crosqmu{v}{(1-A)u}\, =\, -\int_{\ensuremath{\mathbb R}} \int_{\bbS} v(\theta,\omega)\left(\int_0^\theta \frac{{\ensuremath{\mathcal U}} }{q_0} \right)\dd\theta\dd\mu+\frac12 \int_{\ensuremath{\mathbb R}} \int_{\bbS} \frac{vu}{q_0}\dd\theta\dd\mu \\
-\int_{{\ensuremath{\mathbb R}} }\int_{({\ensuremath{\mathbb S}} )^2} v(\theta,\omega)\tilde J*u(\theta,\omega')\dd\theta \dd\mu \otimes\mu \, .
\end{multline}
The right side of this expression is still defined for $u,v\in L^2(\lambda\otimes\mu)$ (recall that $\lambda$ denotes the Lebesgue measure
on ${\ensuremath{\mathbb S}} $, and that we denote the usual scalar product on $L^2(\lambda\otimes\mu)$ by $\Vert\cdot\Vert_{2,\mu}$) and there exists
$c>0$ such that
\begin{equation}
\label{ineq:1-A norm vs L2}
\crosqmu{v}{(1-A)u}\, \leq\, c \Vert u\Vert_{2,\mu} \Vert v\Vert_{2,\mu}\, ,
\end{equation}
Furthermore from \eqref{low bound Dirichlet} and Lemma \ref{lem:comparaison norm} we have
\begin{equation}
\label{ineq:1-A norm vs L2 bis}
\crosqmu{u}{(1-A)u}\, \geq\, \frac1c \Vert u\Vert_{2,\mu}^2\, .
\end{equation}
So the bilinear form $(u,v)\mapsto \crosqmu{v}{(1-A)u}$ is continuous and coercive on $H^{-1}_{\mu}\times H^{-1}_{\mu}$. If
$f\in H^{-1}_{\mu}$, the linear form $v\mapsto \crosqmu{v}{f} $ is continuous on $L^2(\lambda\otimes\mu)$, therefore from Lax-Milgram
Theorem we get that there exists a unique $u\in L^2(\lambda\otimes\mu)$ such that for all $v\in L^2(\lambda\otimes\mu)$
\begin{equation}
\crosqmu{v}{(1-A)u} \, =\, \crosqmu{v}{f} \, .
\end{equation}
Since
\begin{equation}
\crosqmu{v}{f} \, =\, -\int_{\ensuremath{\mathbb R}} \int_{\bbS} v(\theta,\omega)\left(\int_0^\theta \frac{{\ensuremath{\mathcal F}} }{q_0} \right)\dd\theta\dd\mu\, ,
\end{equation}
from \eqref{expr scal prod 1-A} we obtain that for almost $\theta$ and $\omega$
\begin{equation}
-\int_0^\theta \frac{{\ensuremath{\mathcal U}} (\theta',\omega)}{q_0(\theta')}\dd\theta' +\frac{u(\theta,\omega)}{2q_0(\theta)} -\int_{\ensuremath{\mathbb R}} \left(\tilde
J*u\right)(\theta,\omega)\dd\mu\, =\, -\int_0^\theta\frac{{\ensuremath{\mathcal F}} (\theta',\omega)}{q_0(\theta')}\dd\theta'\, .
\end{equation}
So it is clear that if $f$ is continuous with respect to $\theta$, then $u$ has a version $C^2$ with respect to $\theta$. Thus $u\in
D(A)$ and applying $\partial_\theta(q_0(\theta)\partial_\theta \cdot)$ to the both sides of this last expression, we get $(1-A)u=f$.
Since this kind of functions $f$ is dense in $H^{-1}_{\mu}$, we can conclude that the range of $1-A$ is dense, and that $A$ is
essentially self-adjoint.
This completes the proof of Proposition~\ref{th:spectral gap A}.
\hfill $\quad \Box$ \bigskip
\section{Perturbation arguments (completion of the proof of Theorem~\ref{th:expansion speed})}
\label{sec:pa}
In this section we complete the proof of
Theorem \ref{th:expansion speed}.
Essentially, this section is devoted to computing
the expansion of the speed $c_\mu(\delta)$ in. We first recall a lemma that gives a useful parametrization
in the neighborhood of $M_0$. The proof of this lemma is given in \cite{cf:SellYou} , and it is used in the proof of Theorem \ref{th:M} (see \cite{cf:GPPP,cf:SellYou}).
\begin{lemma}
\label{lem:parametrisation}
There exists a $\sigma>0$ such that for all $p$ in the neighborhood
\begin{equation}
\label{eq:Nsigma}
N_\sigma\, :=\,
\cup_{q\in M_0} B_{L^2(\lambda\otimes\mu)}(q,\sigma)\, ,
\end{equation}
of $M_0$ there is one
and only one $q=v(p) \in M_0$ such that $\crosqmu{p-q}{\partial_\theta q}=0$. Furthermore the mapping $p \mapsto v(p)$ is in
$C^\infty(X^1_\mu,X^1_\mu)$, and
\begin{equation}
Dv(p)\, =\, P^o_{v(p)}\, .
\end{equation}
\end{lemma}
\medskip
\noindent
{\it Proof of Theorem \ref{th:expansion speed}.}
The existence and stability
of a {\sl rotating} solution $\tilde q_\delta(\theta-\psi-c_\mu(\delta)t)$ of \eqref{FKP kuramoto disorder delta} ($\psi$ is arbitrary)
has been established in Section~\ref{sec:hyperbolic structure} for $\delta\leq \delta_0$,
see Theorem~\ref{th:M} and the two remarks that follow it.
We are left with proving Theorem \ref{th:expansion speed}(2).
Thanks to the invariance by rotation, we can define $\tilde q_\delta$ such that $v(\tilde q_\delta)=q_0 $.
Now if we denote
\begin{equation}
n_\delta\, :=\, \tilde q_\delta-v\left(\tilde q_\delta\right)\, ,
\end{equation}
then $n_\delta$ verifies $n_\delta=\phi_\delta(q_0)$ and (see Lemma \ref{lem:parametrisation})
\begin{equation}
\label{eq:orth n q}
\crosqmu{n_\delta}{q^\prime_0}=0
\end{equation}
\begin{equation}
\label{eq:orth An q}
\crosqmu{An_\delta}{q^\prime_0}=0\, .
\end{equation}
Moreover the estimates we have on the mapping $\phi_\delta$ in Theorem \ref{th:M} give
\begin{equation}
\label{ineq:bound n}
\Vert n_\delta \Vert_{2,\mu}\, \leq\, C\delta\, ,
\end{equation}
\begin{equation}
\label{ineq:bound dn}
\Vert \partial_\theta n_\delta \Vert_{2,\mu}\, \leq C \delta\, .
\end{equation}
Taking the derivative with respect to $t$, at time $t=0$, we get (we recall the notation $p^{(\psi)}_t(\theta,\omega)=\tilde q_\delta(\theta-\psi-c_\mu(\delta)t)$) :
\begin{equation}
\label{eq:equality derivate}
-c_\mu(\delta)\left(q^\prime_0+\partial_\theta n_\delta\right)\, =\, \partial_t p^{(0)}_0\, .
\end{equation}
So \eqref{FKP kuramoto disorder delta} at time $t=0$ becomes (recall that $q_0$ is a stationary solution of \eqref{FKP kuramoto disorder without drift}) :
\begin{equation}
\label{eq:FKP delta projected}
-c_\mu(\delta)\left( q^\prime_0+\partial_\theta n_\delta\right)\, =\, A n_\delta -\partial_\theta\left[n_\delta\langle J*n_\delta\rangle_\mu\right]-\delta\omega q^\prime_0-\delta\omega\partial_\theta n_\delta\, .
\end{equation}
From \eqref{ineq:bound n} we deduce the bound
\begin{equation}
\label{ineq:bound nJn}
\left\Vert \partial_\theta\left[n_\delta\langle J*n_\delta\rangle_\mu\right]\right\Vert_{-1,\mu}\, \leq \, \Vert J\Vert_2 C^2
\delta^2\, ,
\end{equation}
so by taking the $H^{-1}_{{q_0},\mu}$ scalar product of $q^\prime$ in \eqref{eq:FKP delta projected}, using \eqref{eq:orth An q},
\eqref{ineq:bound n}, \eqref{ineq:bound dn} and the fact that $\int_{\ensuremath{\mathbb R}} w\dd\mu=0$, we get that $c_\mu(\delta)$ is of order $\delta^2$. This implies,
using the same arguments, that
\begin{equation}
\left\Vert A n_\delta-\delta\omega q^\prime_0 \right\Vert_{-1,\mu}\, =\, O(\delta^2)\, .
\end{equation}
So
\begin{equation}
\label{comparaison n0 n1}
\Vert A(n_\delta-\delta n^{(1)}) \Vert_{-1,\mu}\, =\, O(\delta^2)\, ,
\end{equation}
and since $\Vert (1-A)^{(1/2)}u\Vert_{-1,\mu} \sim \Vert u \Vert_{2,\mu}$ (see \eqref{ineq:1-A norm vs L2} and \eqref{ineq:1-A norm vs L2 bis}), we have in particular
\begin{equation}
\Vert n_\delta-\delta n^{(1)} \Vert_{2,\mu}\, =\, O(\delta^2)\, .
\end{equation}
It allows us to make a second order expansion for $c_\mu(\delta)$ : taking again the $H^{-1}_{{q_0},\mu}$ scalar product of $q^\prime_0$ in
\eqref{eq:FKP delta projected}, using the same bounds as for the first order expansion and \eqref{comparaison n0 n1}, we get :
\begin{equation}
\label{eq:second order expansion psi prime}
c_\mu(\delta)\, =\, \delta^2 \frac{\crosqmu{\omega\partial_\theta n^{(1)}+n^{(1)}\langle J*n^{(1)}\rangle_\mu}{q^\prime_0}}{\crosqmu{q^\prime_0}{q^\prime_0}} +O(\delta^3)\, .
\end{equation}
Indeed, from \eqref{comparaison n0 n1}, $\Vert \omega\partial_\theta(n_\delta-\delta n^{(1)}) \Vert_{-1,\mu}$, $\Vert
\partial_\theta[(n_\delta-\delta n^{(1)})\langle J*n^{(1)}\rangle_\mu]\Vert_{-1,\mu}$, $ \Vert \partial_\theta[n^{(1)}\langle J*(n_\delta-\delta
n^{(1)})\rangle_\mu]\Vert_{-1,\mu}$ are of order $\delta^2$ and $ \Vert \partial_\theta[(n_\delta-\delta n^{(1)})\langle J*(n_\delta-\delta
n^{(1)})\rangle_\mu]\Vert_{-1,\mu}$ of order $\delta^4$.
Since $c_\mu(\delta)$ is odd with respect to $\delta$, the second order term in \eqref{eq:second order expansion psi prime} is equal
to $0$. It is possible to get this fact directly : we remark that $n^{(1)}$ satisfies :
\begin{equation}
\label{eq:kdge3}
L_{q_0}\int_{\ensuremath{\mathbb R}} n^{(1)}\dd\mu \, =\, \int_{\ensuremath{\mathbb R}} A n^{(1)} \dd\mu\, =\, \left(\int_{\ensuremath{\mathbb R}} \omega\dd\mu\right)q^\prime_0\, =\, 0\, ,
\end{equation}
\begin{equation}
\crosq{\int_{\ensuremath{\mathbb R}} n^{(1)}\dd\mu}{q^\prime_0}\, =\, \crosqmu{n^{(1)}}{q^\prime_0}\, =\, 0\, .
\end{equation}
So since $L_{q_0}$ is bijective on the orthogonal of $q^\prime_0$ in $H^{-1}_{1/q}$ (see \cite{cf:BGP}), we have $\int_{\ensuremath{\mathbb R}} n^{(1)} \dd\mu=0$ and
$\langle J*n^{(1)}\rangle\mu=0$. On the other hand, since the operator $A$ conserves the parity with respect to $\theta$, $n^{(1)}$ is odd with respect to $\theta$ and thus
\begin{equation}
\crosqmu{\omega\partial_\theta n^{(1)}}{q^\prime_0}\, =\, \int_{\ensuremath{\mathbb S}} \int_{\ensuremath{\mathbb R}} \frac{\omega n^{(1)}}{q_0}\left(q_0-\frac{1}{2\pi I_0^2(2Kr_0)}\right)\dd\theta\dd\mu\, =\, 0\, .
\end{equation}
Now back to \eqref{eq:FKP delta projected}: since $c_\mu(\delta)$ is of order $\delta^3$ and using $\int_{\ensuremath{\mathbb S}} n^{(1)}\dd\mu=0$, we get
\begin{equation}
\left\Vert A \left(n_\delta-\delta n^{(1)}-\delta^2\omega\partial_\theta n^{(1)} \right)\right\Vert_{-1,\mu}\, =\, O(\delta^3)\, ,
\end{equation}
and thus
\begin{equation}
\label{eq:n_2}
\Vert n_\delta-\delta n^{(1)}-\delta^2 n^{(2)}\Vert_{2,\mu} \, =\, O(\delta^3)\, .
\end{equation}
This allows us this time to do a third order expansion in \eqref{eq:FKP delta projected} :
\begin{equation}
\label{eq:rt5p}
c_\mu(\delta)\, =\, \delta^3\frac{\crosqmu{\omega\partial_\theta n^{(2)}}{q^\prime_0}}{\crosqmu{q^\prime_0}{q^\prime_0}} + O(\delta^4)\, .
\end{equation}
This procedure may be repeated recursively at any order: we do not go through the
details again, but we do report the result below (Remark~\ref{rem:pa}) and
we
point out that
the $O(\delta^4)$ \eqref{eq:rt5p} turns out to be $O(\delta^5)$,
in agreement with the fact that $c_\mu(\delta)$ is odd in $\delta$.
\hfill $\quad \Box$ \bigskip
\medskip
\begin{rem}
\label{rem:pa}
\rm
As anticipated above, one can get arbitrarily many terms
in the formal series $c_\mu(\delta)= \sum_{i=1,2, \ldots} c_{2i+1} \delta^{2i+1}$
and the remainder, when the series is stopped at $i=n$, is $O(\delta^{2i+3})$.
In fact, by arguing like above, we have
\begin{equation}
c_5\, =\, \frac{\left\langle \partial_\theta [n^{(2)}\langle J*n^{(3)}\rangle_\mu]+ \partial_\theta [n^{(3)}\langle J*n^{(2)}\rangle_\mu]+ w\partial_\theta n^{(4)} ,q'_0\right\rangle_{-1,q_0,\mu}}{\langle q'_0,q'_0 \rangle_{-1,q_0}}\, ,
\end{equation}
where
\begin{equation}
An^{(3)}\, =\, \partial_\theta [n^{(1)}\langle J*n^{(2)}\rangle_\mu]+ w\partial_\theta n^{(2)}-\frac{\left\langle w\partial_\theta n^{(2)} ,q'_0\right\rangle_{-1,q_0,\mu}}{\langle q'_0,q'_0 \rangle_{-1,q_0}}q'_0 \, ,
\end{equation}
and
\begin{equation}
An^{(4)}\, =\, \partial_\theta [n^{(2)}\langle J*n^{(2)}\rangle_\mu]+ \partial_\theta [n^{(1)}\langle J*n^{(3)}\rangle_\mu]+ w\partial_\theta n^{(3)}\, .
\end{equation}
Actually,
by induction we obtain
\begin{equation}
c_{2i+1}\, =\, \frac{\left\langle \sum_{k+l=2i+1,k>0,l>0}\partial_\theta [n^{(l)}\langle J*n^{(k)}\rangle_\mu]+ w\partial_\theta n^{(2i)} ,q'_0\right\rangle_{-1,q_0,\mu}}{\langle q'_0,q'_0 \rangle_{-1,q_0}}\, ,
\end{equation}
and
\begin{equation}
n^{(2i)}\, =\, \sum_{k+l=2i,k>0,l>0}\partial_\theta [n^{(l)}\langle J*n^{(k)}\rangle_\mu] + w\partial_\theta n^{(2i-1)}\, ,
\end{equation}
\begin{equation}
n^{(2i+1)}\, =\, \sum_{k+l=2i+1,k>0,l>0}\partial_\theta [n^{(l)}\langle J*n^{(k)}\rangle_\mu] + w\partial_\theta n^{(2i)}- c_{2i+1}q'_0\, .
\end{equation}
Since this procedure yields also $n^{(j)}$ for arbitrary $j$, one
can generalizes also \eqref{eq:n_2} and, hence, \eqref{eq:qgddevel}.
\end{rem}
\section{Active rotators}
\label{sec:AR}
In this section we deal with the equation \eqref{eq:AR} and we do it in a rather informal way, because on one hand a formal statement would be very close to
Theorem~\ref{th:expansion speed} and, on the the other hand, the large scale behavior
of disordered active rotators is qualitatively and quantitatively close to the non disordered case, treated in
\cite{cf:GPPP}, in a way that we explain below.
First of all, from a technical viewpoint the main difference between
\eqref{eq:AR} and \eqref{FKP kuramoto disorder} is that \eqref{eq:AR}
is (in general) not rotation invariant, so the manifold $M_\delta=\{q_\psi+\phi(q_\psi)\}$ we get after perturbation is not necessarily a circle. Unlike Theorem \ref{th:expansion speed}, the motion on $M_\delta$ is not uniform, and we describe the behaviour on $M_\delta$ by the phase derivate $\dot{\psi}$. We follow the same procedure as in the previous section : if $p^\delta_t$ is a solution \eqref{eq:AR} belonging to $M_\delta$, we define (see Lemma \ref{lem:parametrisation})
\begin{equation}
q_{\psi^\delta_t}\, =\, v(p^\delta_t)\ , \ \ \text{ and } \ \ \
n^\delta_t\,=\, p^\delta_t- v(p^\delta_t)\, .
\end{equation}
In this context, \eqref{eq:FKP delta projected} becomes
\begin{equation}
\label{eq:FKP delta projected AR}
-\dot{\psi}^\delta_t q'_{\psi^\delta_t} + \partial_t n^\delta_t\, =\, A^{\psi^\delta_t} n^\delta_t -\partial_\theta [n^\delta_t \langle J*n^\delta_t \rangle_\mu ]-\delta U q'_\psi -\delta U\partial_\theta n^\delta_t \, ,
\end{equation}
where $A^\psi$ is the rotation of the operator $A$
\begin{equation}
A^\psi u(\theta,\omega) \, :=\, \frac12 \Delta u(\theta,\omega)-\partial_\theta\Big(q_0(\theta-\psi)\langle J*u\rangle_\mu(\theta)+u(\theta,\omega)J*q_0(\theta-\psi)\Big)\, .
\end{equation}
Note that we can reformulate the second term of the left hand side in \eqref{eq:FKP delta projected AR}:
\begin{equation}
\partial_t n^\delta_t\, =\, \dot{\psi}^\delta_t\partial_\psi \phi(q_\psi) |_{\psi=\psi^\delta_t}\, .
\end{equation}
So, as in the previous section, using the estimates on the mapping $\phi$ given in Theorem \ref{th:M}, we get the bounds
\begin{equation}
\Vert n^\delta_t \Vert_{2,\mu}\, \leq\, C\delta \, , \ \
\Vert \partial_t n^\delta_t \Vert_{2,\mu}\, \leq\, C\delta |\dot{\psi}^\delta_t|
\ \text{ and } \
\Vert \partial_\theta [n^\delta_t \langle J*n^\delta_t \rangle_\mu ] \Vert_{2,\mu} \, \leq\, \Vert J\Vert_2 C^2\delta \, ,
\end{equation}
and we deduce the first order expansion
\begin{equation}
\label{eq:AReq}
{\dot{\psi}}^\delta_t\,
=\, \delta\frac{ \langle (U q_{\psi^\delta_t})', q'_{\psi^\delta_t} \rangle_{-1,q_{\psi^\delta_t},\mu}}{ \langle q'_0,q'_0 \rangle_{-1,q_0}}+O(\delta^2)\, .
\end{equation}
Since $\dot{\psi}$ is odd in $\delta$ and the expansion can be pushed further in $\delta$, this $O(\delta^2)$ is in reality a $O(\delta^3)$ and one can actually improve this result both in the direction
of obtaining a regularity estimate on the $O(\delta^2)$ rest in \eqref{eq:AReq}
(like in \cite[Th.~2.3]{cf:GPPP}) and of going to higher orders (like in Remark~\ref{rem:pa}).
However the evolution for small $\delta$ is dominated by the leading order and from
\eqref{eq:AReq} we can directly read that, to first order, the effect of the disorder
is rather simple: in fact
\begin{equation}
\langle (U q_{\psi})', q'_{\psi} \rangle_{-1,q_{\psi},\mu}\, =\,
\int_{\ensuremath{\mathbb R}} \int_{\ensuremath{\mathbb S}} U(\theta, \omega) q_{\psi} (\theta)\left( q_{\psi}(\theta)- c\right) \dd
\theta \mu (\dd \omega)\, ,
\end{equation}
where $c$ is such that $\int_{\ensuremath{\mathbb S}} (q_\psi -c)=0$, that is $1/c={2\pi (I_0(2Kr_0))^2}$
(recall \eqref{eq:defstationarysolution nodisorder}-\eqref{def bessel}: this computation
is analogous to \eqref{eq:kdge3}).
Since the integrand depends on $\omega$ only via $U$, this integration can be performed first
and the system behaves to leading order in $\delta$ as the non-disordered model
with active rotator dynamics led by the deterministic force
$\int_{\ensuremath{\mathbb R}} U(\cdot, \omega) \mu(\dd \omega)$. The rich phenomenology connected
to these models is worked out in \cite[Sec.~3]{cf:GPPP}.
\section{Symmetric case: stability of the stationary solutions}
\label{sec:sym}
\subsection{On the non-trivial stationary solutions (proof of Lemma \ref{th: Psimu concave})}
We start by observing that
in the case with no disorder the strict concavity of the fixed-point function $\Psi_0$ has been proven in \cite[Lemma 4, p.315]{cf:Pearce}, in
the apparently different context of classical XY-spin model (for a detailed discussion on the link with these models see \cite{cf:BGP}). We are
going to obtain the concavity of $\Psi^\mu_\delta$ for small $\delta$ via a perturbation argument, by relying on the result in \cite{cf:Pearce}.
Since $\Psi^\mu_\delta$ is a smooth perturbation of $\Psi_0$, one expects that the strict concavity of $\Psi_0$ will be preserved
to $\Psi^\mu_\delta$ for small $\delta>0$, namely $\sup_{x} (\Psi_\delta^\mu)''(x)<0$. Nevertheless, an easy calculation shows that
$\Psi''_0(0)=0$; in that sense one has to treat the concavity in a neighborhood of $0$ as a special case.
\medskip
\noindent
In what follows, we suppose that the coupling strength $K$ is bounded above and below by fixed constants $K_{\min}$ and $K_{\max}$:
\begin{equation}0\, <\, K_{\min} \, \leq\, K\, \leq\, K_{\max}\, <\, \infty\, .\end{equation}
We first prove the statement on the concavity in a neighborhood of $0$: there exist $\eta_0>0$, $\delta>0$ such
that for all $K\in[K_{\min}, K_{\max}]$, for all $\mu$ such that $\Supp(\mu)\subseteq[-1,1]$, $\Psi_\delta^\mu$ is strictly concave on $[0,
\eta_0]$.
\medskip
Indeed, one easily shows (using that the function $x\mapsto \Psi_\mu^\delta(x)$ is odd) that we have the following Taylor's expansion:
\begin{equation}(\Psi_\delta^\mu)''(x) = -6 D^\delta(\mu)K^3 x + \epsilon(x)\, ,\end{equation}where $\epsilon(x)=o(x)$ as $x\rightarrow 0$ and where
for fixed $\mu$, we write \begin{equation}D^\delta(\mu) := \int_{\mathbb{R}{}}{h(\delta\omega)\mu(\dd\omega)},\end{equation}where \begin{equation}h(\omega)
:= \frac{1}{2(1+\omega^{2})}-\frac{8\omega^{2}}{(1+4\omega^{2})^2}.\end{equation}
Note that the $o(x)$ only depends on $K_{\max}$ (in particular it can be chosen independently of $\mu$).
A closer look at the function $h$ shows that there exists $\delta>0$ such that for all $\mu$ with $\Supp(\mu)\subseteq[-1, 1]$, $D^\delta(\mu)>\frac
14$.
If we choose $\eta_{0}>0$ such that $\frac{1}{\eta_0}\sup\limits_{0\leq x< \eta_0}|\epsilon(x)|< \frac 32 K_{\min}^3$ then
$(\Psi_\delta^\mu)''(x)<0$ for all $0<x<\eta_0$, which is the desired result.
\medskip
We are now left with proving concavity away from $0$: namely, we prove that for all $\eta>0$, all $K_{\max}$, there exists $\delta_0>0$ such that
for all $K\leq K_{\max}$, for all $0<\delta<\delta_0$, for any measure $\mu$ such that $\Supp(\mu)\subseteq[-1, 1]$, $\Psi^\mu_\delta$ is strictly concave
on $[\eta, 2K_{\max}]$.
\medskip
Indeed, using the strict concavity of $\Psi_0$ proved in \cite{cf:Pearce}, there exists a constant $\alpha>0$ such that for all
$x\in[\eta, 2K_{\max}]$, $\Psi_{0}''(x)<-\alpha<0$. But then, it easy to see that \begin{equation}\sup_{0<\delta<\delta_0}\sup_{\mu,\
\Supp(\mu)\subseteq [-1,
1]}\sup_{x\in[0, 2K_{\max}]} \left|(\Psi^\mu_\delta)''(x) - \Psi_{0}''(x)\right|
\stackrel{\delta_0\searrow 0}{\rightarrow}0.\end{equation}If one chooses
$\delta_0$ such that the latter quantity is smaller than or equal to $\frac{\alpha}{2}$, the result follows. The proof of Lemma \ref{th: Psimu concave} is therefore complete.
\hfill $\quad \Box$ \bigskip
\subsection{On the linear stability of non-trivial stationary solutions}
We now prove Theorem \ref{th:spectral prop L disorder} along with a number of explicit estimates.
\begin{rem}\rm
Note that, since the whole operator $L^\omega_{{q}}$ is no longer self-adjoint nor symmetric, its spectrum need not be real. In that
extent, one has to deal in this section with the complexified versions of the scalar products defined in Section \ref{sec:mainresults},
\eqref{def scalar product sobolev with wage and mu} and in Section \ref{sec:hyperbolic structure}, \eqref{def scalar product L2 with wage and
mu}. Thus, we will assume for the rest of this section that we work with complex versions of these scalar products. The results
concerning the operator $A$ are obviously still valid, since $A$ is symmetric and real.
We will also use the following standard notations: for an operator $F$, we will denote by $\rho(F)$ the set of all complex numbers $\lambda$ for
which $\lambda-F$ is invertible, and by $R(\lambda, F):= \left( \lambda - F \right)^{-1}$, $\lambda\in\rho(F)$ the resolvent of $F$.
The spectrum of $F$ will be denoted as $\sigma(F)$.
\end{rem}
\subsubsection{Decomposition of $L^\omega_{{q}}$}
In what follows, $K>1$ and $r_0= \Psi_{0}(2Kr_0)>0$ are fixed.
In order to study the spectral properties of the operator $L^\omega_{{q}}$ for general distribution of disorder, we decompose
$L^\omega_{{q}}$ in
\eqref{eq:defLqmu} into the sum of the self-adjoint operator $A$ defined in \eqref{def A} and a perturbation $B$ which
will be considered to be small w.r.t.
$A$, namely:
\begin{equation}
\label{eq:defBmu}
Bu(\theta, \omega)\, :=\, - \partial_{\theta} \left( u(\theta, \omega) \langle J \ast \gep(q) \rangle_\mu + \gep(q)(\theta, \omega, \delta)
\langle J \ast
u\rangle_\mu(\theta) + \delta\omega u(\theta, \omega)\right),
\end{equation}
where
\begin{equation}
\label{eq:defdeltaqmu}
\gep(q) \, :=\, (\theta, \omega, \delta)\mapsto q(\theta, \delta\omega) - q_0(\theta),
\end{equation}
is the difference between the stationary solution with disorder and the one without disorder.
\begin{proposition}
\label{prop:Apositivemu}
The (extension of the) operator $A$ is the infinitesimal generator of a strongly continuous semi-group
of contractions $T_A(t)$ on $H^{-1}_{{q_0},\mu}$.
Moreover, for every $0<\alpha<\frac{\pi}{2}$ this semigroup can be extended to an
analytic semigroup $T_A(z)$ defined on $\Delta_\alpha\, :=\, \ens{z\in{\ensuremath{\mathbb C}} }{|\arg(z)|<\alpha}$.
\end{proposition}
We recall here the result we use concerning analytic extensions of strongly continuous semigroups. Its proof can be found in
\cite[Th 5.2, p.61]{Pazy1983}.
\begin{proposition}
\label{prop:pazysemgps}
Let $T(t)$ a uniformly bounded strongly continuous semigroup, whose infinitesimal generator $F$ is such that $0\in\rho(F)$ and let
$\alpha\in(0, \frac\pi2)$. The
following statements are equivalent:
\begin{enumerate}
\item \label{it:prop:pazysemgps1}$T(t)$ can be extended to an analytic semigroup
in the sector $\Delta_\alpha\, =\, \ens{\lambda\in{\ensuremath{\mathbf C}} }{|\arg(\lambda)|<\alpha}$ and $\Vert{T(z)}\Vert$ is uniformly bounded in every closed
sub-sector $\bar{\Delta}_\alpha'$, $\alpha'<\alpha$, of $\Delta_\alpha$,
\item \label{it:prop:pazysemgps2} There exists $M>0$ such that\begin{equation}\rho(F) \supset \Sigma\, =\,
\ens{\lambda
\in{\ensuremath{\mathbf C}} }{|\arg(\lambda)|<\frac{\pi}{2}+\alpha} \cup \{0\},\end{equation}and\begin{equation}\Vert{R(\lambda, F)}\Vert \, \leq\, \frac{M}{|\lambda|}, \quad
\lambda\in\Sigma, \lambda\neq0\, .\end{equation}
\end{enumerate}
\end{proposition}
\begin{proof}[Proof of Proposition \ref{prop:Apositivemu}]
The proof in Section \ref{sec:hyperbolic structure}, Theorem \ref{th:spectral gap A} of the self-adjointness of $A$ shows that $A$ satisfies
the hypothesis of Lumer-Phillips Theorem (see \cite[Th 4.3, p.14]{Pazy1983}): $A$ is the infinitesimal generator of a $C_0$ semi-group of
contractions denoted by $T_A(t)$.
The rest of the proof is devoted to show the existence of an analytic extension of this semigroup in a proper sector. We follow
here the lines of the proof of Th 5.2, p. 61-62, in \cite{Pazy1983}, but with explicit estimates on the resolvent, in order to quantify properly
the appropriate size of the perturbation.
Let us first replace the operator $A$ by a small perturbation: for all $\gep>0$, let $A_{\gep}\, :=\, A-\gep$, so that
$0$ belongs to $\rho(A_\gep)$. The operator $A_{\gep}$ has the following
properties: as $A$, it generates a strongly continuous semigroup of operators (which is $T_{A, \gep}(t)=T_A(t)e^{-\gep t}$).
Since $A$ is self-adjoint, it is easy to see that
\begin{equation}
\label{eq:estimRAeps1}
\forall
\lambda\in{\ensuremath{\mathbf C}} \smallsetminus{\ensuremath{\mathbb R}} , \Nsqmu{R(\lambda, A_\gep)}\, \leq\,
\frac{1}{|\Im(\lambda)|}\, ,
\end{equation}
and since the spectrum of $A$ is negative, for every $\lambda\in{\ensuremath{\mathbf C}} $ such that $\Re(\lambda)>0$,
\begin{equation}
\label{eq:estimRAepsdef1}
\Nsqmu{R(\lambda, A_\gep)}\, \leq\, \frac{1}{|\lambda|}\, .
\end{equation}
For any $\alpha \in(0, \frac{\pi}{2})$, let
\begin{equation}
\Sigma_\alpha \,:=\, \ens{\lambda\in{\ensuremath{\mathbb C}} }{|\arg(\lambda)|<\frac{\pi}{2}+\alpha}\, .
\end{equation}
Let us prove that for $\lambda\in\Sigma_\alpha$,
\begin{equation}
\label{eq:estim R lambda alpha}
\Nsqmu{R(\lambda, A_\gep)}\, \leq\, \frac{1}{1-\sin(\alpha)}\cdot \frac{1}{|\lambda|}\, .
\end{equation}
Note that \eqref{eq:estim R lambda alpha} is clear from \eqref{eq:estimRAeps1} and \eqref{eq:estimRAepsdef1} when $\lambda$ is such
that $\Re(\lambda)\geq0$.
Let us consider $\sigma>0, \tau\in{\ensuremath{\mathbb R}} $ to be chosen appropriately later.
Let us write the following Taylor expansion for
$R(\lambda, A_\gep)$ around $\sigma+i\tau$ (at least well defined in a neighborhood of $\sigma+i\tau$ since $\sigma>0$):
\begin{equation}
\label{eq:taylor resolvent}
R(\lambda, A_\gep) \, =\, \sum_{n=0}^{\infty}{R(\sigma+i\tau, A_\gep)^{n+1}((\sigma + i\tau)-\lambda)^n}\, .
\end{equation}
From now, we fix $\lambda\in\Sigma_\alpha$ with $\Re(\lambda)<0$.
This series $R(\lambda, A_\gep)$ is well defined in $\lambda$ if one can choose $\sigma$, $\tau$ and $k\in(0,1)$ such that $\Nsqmu{R(\sigma
+i\tau, A_\gep)}|\lambda-(\sigma+i\tau)|\leq k<1$. In particular, using \eqref{eq:estimRAeps1}, it suffices to have $|\lambda-(\sigma+i\tau)|\leq
k|\tau|$ and since $\sigma>0$ is
arbitrary, it suffices to find $k\in(0,1)$ and $\tau$ with $|\lambda-i\tau|\leq k|\tau|$ to obtain the convergence of \eqref{eq:taylor
resolvent}.
For this $\lambda\in\Sigma_\alpha$ with $\Re(\lambda)<0$, let us define $\lambda'$ and $\tau$ as in Figure~\ref{fig:angle alpha}. Then,
$|\lambda-i\tau|\leq |\lambda'-i\tau|= \sin(\alpha)|\tau|$ with $\sin(\alpha)\in(0,1)$. So
the series converges for $\lambda\in\Sigma_\alpha$ and one has, using again \eqref{eq:estimRAeps1},
\begin{equation}
\label{eq:estimRAepsdef}
\Nsqmu{R(\lambda, A_\gep)} \, \leq\, \frac{1}{(1-\sin(\alpha))|\tau|} \, \leq\, \frac{1}{1-\sin(\alpha)}
\cdot\frac{1}{|\lambda|}\, .
\end{equation}
\begin{figure}[!ht]
\centering
\includegraphics[width=0.5\textwidth]{resolvent.eps}
\caption{The set $\Sigma_\alpha$.}
\label{fig:angle alpha}
\end{figure}
The fact that $T_{A, \gep}(t)$ can be extended to an analytic semigroup $T_{A, \gep}(z)$ on the domain $\Delta_\alpha$ is a simple
application of \eqref{eq:estimRAepsdef} and Proposition \ref{prop:pazysemgps}, with $M:=\frac{1}{1-\sin(\alpha)}$.
Let us then define $\tilde{T_A}(z):= e^{\gep z} T_{A,\gep}(z)$, for $z\in\Delta_\alpha$
so that $\tilde{T_A}$ is an
analytic extension of $T_A$ (an argument of analyticity shows that $\tilde{T_A}$ does not depend on $\gep$).
\end{proof}
\begin{rem}\rm
Note that estimate \eqref{eq:estim R lambda alpha} is also valid in the limit as $\gep\to 0$: for all $\alpha\in(0, \frac\pi2)$,
$\lambda\in\Sigma_\alpha$,
\begin{equation}
\label{eq:estim R lambda alpha without gep}
\Nsqmu{R(\lambda, A)}\, \leq\, \frac{1}{1-\sin(\alpha)}\cdot \frac{1}{|\lambda|}\, .
\end{equation}
\end{rem}
\subsubsection{Spectral properties of $L^\omega_{{q}}=A+B$}
In this part, we show that if the perturbation $B$ is small enough with respect to $A$, one has the same spectral properties for
$L^\omega_{{q}}= A+B$ as for $A$. In this extent, we recall that $\mu$ is of compact support in $[-1, 1]$, and the disorder is
rescaled by $\delta>0$.
\begin{proposition}
\label{prop:BAboundedmu}
\item The operator $B$ is $A$-bounded, in the sense that there exist explicit constants $a_{K, \delta}$ and $b_{K,\delta}$, depending
on $K$ and $\delta$ such that for all $u$ in the domain of (the closure of) $A$
\begin{equation}
\label{eq:BAboundedmu}
\Nsqmu{Bu}\, \leq\, a_{K, \delta}\Nsqmu{u} + b_{K, \delta}\Nsqmu{Au}\, .
\end{equation}
Moreover, for fixed $K>1$, $a_{K,\delta}=O(\delta)$ and $b_{K,\delta}=O(\delta)$, as $\delta\rightarrow0$.
\end{proposition}
The latter proposition is based on the fact that the difference $\gep(q)(\theta, \omega, \delta) = q(\theta, \delta\omega)- q_0(\theta)$
in \eqref{eq:defdeltaqmu} is small if the scale parameter $\delta$ tend to $0$:
\begin{lemma}
\label{lem:estimdeltaq}
For $\delta>0$, let us define
\begin{equation}
\label{eq:def Ninf eps(q)}
\Ninf{\gep(q)}\, :=\, \suptwo{\theta\in{\ensuremath{\mathbb S}} , |\omega|\leq 1}{0<u<\delta} |\gep(q)(\theta, \omega, u)|\, .
\end{equation}
Then for all $K>1$, $\Ninf{\gep(q)} = O(\delta)$, as $\delta\rightarrow 0$. More precisely, for $K>1$, $\delta>0$, the following
inequality holds:
\begin{equation}
\label{eq:estimdeltaq}
\Ninf{\gep(q)}\, \leq\, \gep_{K,\delta}\, ,
\end{equation}
where the constant $\gep_{K, \delta}$ can be chosen explicitly in terms of $K$ and $\delta$:
\begin{equation}
\label{eq:estimdeltaK}
\gep_{K, \delta}\, :=\, \frac{\delta}{\pi} e^{8\pi\delta}\left( 2+ 3e^{4\pi\delta} \right) e^{14K\bar{r}_\delta}\left( 1+ 2\pi
e^{2K\bar{r}_\delta}\right)\, ,
\end{equation}
where we recall that $\bar{r}_\delta= \max\left( r_0, r_\delta\right)$.
\end{lemma}
\begin{proof}[Proof of Lemma \ref{lem:estimdeltaq}]
Recall that the disordered stationary solution $q$ \eqref{eq:qhatom} is
given by
\begin{equation}
q(\theta, \delta\omega) \,:=\, \frac{S(\theta, \delta\omega, 2Kr_\delta)}{Z(\delta\omega, 2Kr_\delta)},
\end{equation}
where $S(\theta, \omega, x)$ is defined in
\eqref{eq:Sq} and that the non-disordered one
\eqref{eq:defstationarysolution
nodisorder} is given by $q_0(\theta)=\frac{S(\theta, 0, 2Kr_0)}{Z(0, 2Kr_0)}= \frac{e^{2Kr_0\cos(\theta)}}{\int_{\bbS}
e^{2Kr_0\cos(\theta)}\dd\theta}$.
Since $q(\theta, \delta\omega)=q(-\theta, -\delta\omega)$, it suffices to consider the case $\delta\omega>0$.
A simple computation shows that
\begin{equation}
\label{eq:lower bound Zomega}
Z(\delta\omega, 2Kr_\delta)\, \geq\, 4\pi^2 e^{-4K r_\delta}e^{-4\pi\delta}\, ,
\end{equation}
and that
\begin{equation}
\label{eq:upper bound S0}
|S(\theta, 0)|\, \leq\, 2\pi e^{4Kr_0}\, .
\end{equation}
Using $|q(\theta, \delta\omega)- q_0(\theta)|\leq \frac{1}{Z(\delta\omega)Z(0)}\left( Z(0)|S(\theta, \delta\omega) - S(\theta, 0)| +
|S(\theta, 0)||Z(0)-Z(\delta\omega)|\right)$, one has to deal with, successively:
\begin{itemize}
\item for fixed $\theta\in{\ensuremath{\mathbb S}} $, $|S(\theta, \delta\omega) - S(\theta, 0)|\leq \delta\cdot \sup_{|\omega|\leq 1}|\frac{\dd}{\dd\omega} S(\theta,
\delta\omega)|$. A long calculation shows that the latter expression $|\frac{\dd}{\dd\omega} S(\theta, \delta\omega)|$ can be bounded above by
$8\pi^2 e^{4Kr_\delta}e^{4\pi\delta}\left( 2+ 3e^{4\pi\delta}\right)$, that is,
\begin{equation}
\label{eq:upper bound diff S}
|S(\theta, \delta\omega) - S(\theta, 0)|\, \leq\, \delta 8\pi^2 e^{4Kr_\delta}e^{4\pi\delta}\left( 2+ 3e^{4\pi\delta}\right)\, .
\end{equation}
\item Using $|Z(\delta\omega) - Z(0)|= \left|\int_{\bbS} (S(\theta, \delta\omega) - S(\theta, 0)) \dd\theta\right|$ and \eqref{eq:upper bound diff S},
one has directly:
\begin{equation}
\label{eq:upper bound diff Z}
|Z(\delta\omega) - Z(0)|\, \leq\, \delta 16\pi^3 e^{4Kr_\delta}e^{4\pi\delta}\left( 2+ 3e^{4\pi\delta}\right)\, .
\end{equation}
\end{itemize}
Putting together \eqref{eq:lower bound Zomega}, \eqref{eq:upper bound S0}, \eqref{eq:upper bound diff S} and \eqref{eq:upper bound
diff Z}, one obtains the result.
\end{proof}
We are now in position to prove the $A$-boundedness of $B$:
\begin{proof}[Proof of Proposition \ref{prop:BAboundedmu}]
$B$ is $A$-bounded: let us fix a $u$ in the domain of the closure of $A$. Then we have $\Nsqmu{Bu}= \NLqmu{{\ensuremath{\mathcal B}} u}$, where ${\ensuremath{\mathcal B}} u$ is
the
appropriate primitive of $Bu$, namely:
\begin{align}
{\ensuremath{\mathcal B}} u(\theta, \omega)&\, :=\, {} -\left( u(\theta, \omega) \langle J \ast \gep(q) \rangle_\mu + \gep(q)(\theta, \omega, \delta) \langle J \ast
u\rangle_\mu(\theta) + \delta\omega u(\theta, \omega)\right)\nonumber\\
&+{} \left( \int_{\bbS} \frac{1}{q_0} \right)^{-1}\left( \int_{\bbS} \frac{u(\theta, \omega) \langle J \ast \gep(q) \rangle_\mu + \gep(q)(\theta,
\omega, \delta) \langle J \ast
u\rangle_\mu(\theta) + \delta\omega u(\theta, \omega)}{q_0(\theta)} \dd\theta\right)\, .\label{eq:primitiveBu}
\end{align}
One can easily shows that there exists a constant $c^{(1)}_{K, \delta}$, depending only on $K>1$ and $\delta>0$ such
that:
\begin{equation}
\label{eq:estim1Bu}
\Nsqmu{Bu} \, \leq\, c^{(1)}_{K, \delta} \NLqmu{u}\, .
\end{equation}
Indeed, an easy calculation shows that $|\langle J\ast \gep(q)\rangle_\mu|\leq 4K\Ninf{\gep(q)}$ and that
\begin{equation}
\begin{split}
|\langle J\ast
u\rangle_\mu(\cdot)|&\, \leq\, K\left( \int_{\bbS} \sin(\cdot-\varphi)^2 q_0(\varphi)\dd\varphi \right)^{\frac12}\NLqmu{u}\\
&\,\leq K\left( \int_{\bbS} q_0(\varphi)\dd\varphi \right)^{\frac12}\NLqmu{u} = K \NLqmu{u}\, .
\end{split}
\end{equation}
So we have for all $\theta, \omega$ (recall that $Z_0$ is
the normalization constant in
\eqref{eq:defstationarysolution nodisorder}):
\begin{equation}
\begin{split}
|{\ensuremath{\mathcal B}} u(\theta, \omega)|\, \leq\, & \left(4 K\Ninf{\gep(q)}+\delta|\omega|\right)|u|+ 2K \Ninf{\gep(q)}\NLqmu{u}\\ &+
Z_{0}^{-1}\left(4 K
\Ninf{\gep(q)} + \delta|\omega|\right) \left( \int_{\bbS} \frac{|u|^2}{q_0} \right)^{\frac 12}\, .
\end{split}
\end{equation}
Hence, inequality \eqref{eq:estim1Bu} is true for the following choice of $c^{(1)}_{K, \delta}$ (recall that $\gep_{K, \delta}$ is defined in
\eqref{eq:estimdeltaK}):
\begin{equation}
\label{eq:defc1}
c^{(1)}_{K, \delta}\, :=\, \left( 6\left(4 K\gep_{K, \delta}+\delta\right)^2 +12 K^2 Z_{0}^2 \gep_{K, \delta}^2 \right)^{\frac
12}\, .
\end{equation}
\begin{rem}\rm
Note that, thanks to Lemma \ref{lem:estimdeltaq}, one has that $c^{(1)}_{K, \delta}=O(\delta)$ as $\delta\rightarrow0$.
\end{rem}
In order to complete the proof of the inequality \eqref{eq:BAboundedmu}, it suffices to prove that there exist constants
$c^{(2)}_K$
and $c^{(3)}_K$, only depending on $K$ such that, for all $u$:
\begin{equation}
\label{eq:estim2Bu}
\NLqmu{u} \, \leq\, c^{(2)}_K \Nsqmu{Au} + c^{(3)}_K \Nsqmu{u}\, .
\end{equation}
The rest of this first of the proof is devoted to find explicit expressions of $c^{(2)}_K$ and $c^{(3)}_K$, and is based on an
interpolation argument.
For all integer $n>1$, one can compute the linear operator $f\mapsto f'$ in terms of a sum of two integral operators,
namely:
\begin{equation}
\label{eq:identuprimemu}
f' \, =\, \ensuremath{I_n}(f'') + \ensuremath{J_n}(f)\, ,
\end{equation}
where $\ensuremath{I_n}:f\mapsto \int_{0}^{2\pi} i_n(\theta, \varphi)f(\varphi)\dd\varphi$ (resp.
$\ensuremath{J_n}:f\mapsto \int_{0}^{2\pi} j_n(\theta, \varphi)f(\varphi)\dd\varphi$) is the integral operator whose kernel $i_n(\theta,
\varphi)$ (resp. $j_n(\theta, \varphi)$) is defined by:
\begin{equation}\left\{\begin{array}{lll}
i_n(\theta, \varphi)\, :=\, \frac{\varphi^{n+1}}{2\pi\theta^{n}}\, ,& j_n(\theta, \varphi)\, :=\, -\frac{n(n+1)\varphi^{n-1}}{2\pi\theta^n}\, ,&
0\, \leq\, \varphi\, <\, \theta\, \leq\, 2\pi\, ,\\[10pt]
i_n(\theta, \varphi)\, :=\, \frac{-(2\pi-\varphi)^{n+1}}{2\pi(2\pi-\theta)^n}\ , & j_n(\theta, \varphi)\, :=\, \frac{n(n+1)
(2\pi-\varphi)^{n-1}}{2\pi (2\pi-\theta)^n}\ , & 0\, \leq\, \theta\, <\, \varphi\, \leq\, 2\pi\, .
\end{array}\right.\end{equation}
Equality \eqref{eq:identuprimemu} can be easily verified by integrations by parts. Since,
\begin{equation}\left\{\begin{array}{ll}
\int_{0}^{2\pi} \left|i_n(\theta, \varphi)\right|\dd\varphi \, \leq\, \frac{2\pi}{n+2},& \int_{0}^{2\pi} \left|i_n(\theta,
\varphi)\right|\dd\theta\, \leq\, \frac{2\pi}{n-1}\, ,\\[10pt]
\int_{0}^{2\pi} \left|j_n(\theta, \varphi)\right|\dd\varphi \, \leq\, \frac{n+1}{\pi}\, ,& \int_{0}^{2\pi} \left|j_n(\theta,
\varphi)\right|\dd\theta\, \leq\, \frac{n(n+1)}{\pi(n-1)}\, ,
\end{array}\right.\end{equation}
we see (cf. \cite[p.143-144]{Kato1995}) that $\ensuremath{I_n}$ and $\ensuremath{J_n}$ are bounded operators on $L^2({{\ensuremath{\mathbb S}} })$, namely:
\begin{equation}\Vert{\ensuremath{I_n}}\Vert\, \leq\, \frac{2\pi}{n-1},\quad \Vert{\ensuremath{J_n}}\Vert\, \leq\, \frac{n(n+1)}{\pi(n-1)}\, .\end{equation}
So, applying relation \eqref{eq:identuprimemu} for $f={\ensuremath{\mathcal U}} $ we get, for $\mu$-almost every $\omega$:
\begin{equation}
\left( \int_{\bbS} |u(\theta, \omega)|^2 \dd\theta \right)^{\frac 12} \leq\frac{2\pi}{n-1}\left( \int_{\bbS} |u'(\theta, \omega)|^2 \dd\theta
\right)^{\frac 12} +
\frac{n(n+1)}{\pi(n-1)}\left(\int_{\bbS}|{\ensuremath{\mathcal U}} (\theta, \omega)|^2 \dd\theta \right)^{\frac 12}.
\end{equation}
This gives
\begin{equation}
\label{eq:estuUmu}
\NLmu{u}\, \leq\,\frac{2\pi}{n-1}\NLmu{u'} +\frac{n(n+1)}{\pi(n-1)}\NLmu{{\ensuremath{\mathcal U}} }\, .
\end{equation}
Since $\NLqmu{{\ensuremath{\mathcal U}} }= \Nsqmu{u}$, it only remains to control $\NLqmu{u'}$ with $\Nsqmu{Au}$:
like for the beginning of this proof for the operator $B$, we have $\Nsqmu{Au}= \NLqmu{{\ensuremath{\mathcal A}} u}$, where ${\ensuremath{\mathcal A}} u$ is the
appropriate primitive of $Au$:
\begin{align}
{\ensuremath{\mathcal A}} u(\theta, \omega)&\, :=\, {} {\frac 12} u'(\theta, \omega) -\left( u(\theta, \omega) (J \ast q_0) + q_0(\theta) \langle J
\ast u\rangle_\mu(\theta)\right)\nonumber\\
&+{} \left( \int_{\bbS} \frac{1}{q_0} \right)^{-1}\left( \int_{\bbS} \left\{\frac{u(\theta, \omega) (J \ast
q_0)}{q_0(\theta)} + {\frac 12} u(\theta, \omega) \partial_\theta\left( \frac{1}{q_0(\theta)}
\right)\right\}\dd\theta\right)\, .\label{eq:primitiveAu}
\end{align}
Using inequalities $|\langle J\ast u\rangle|_{\mu}(\cdot)\leq K\sqrt{\pi}\NLmu{u}$, and $\int_{\bbS} \frac{|u(\cdot, \omega)|}{q_0}\leq
Z_0^{\frac12}e^{Kr_0}\left( \int_{\bbS} |u(\cdot, \omega)^2|\right)^{\frac12}$, an easy calculation shows that:
\begin{equation}
\label{eq:estimuprimeAu1}
|u'(\cdot, \omega)|\, \leq\, 2|{\ensuremath{\mathcal A}} u(\cdot, \omega)| + 2Kr_0|u(\cdot, \omega)|+2\sqrt{\pi}K q_0(\cdot)\NLmu{u} +
\frac{4Kr_0}{Z_0^\frac12}e^{Kr_0}\left( \int_{\bbS} |u(\cdot, \omega)^2|\right)^{\frac12}\, ,
\end{equation}
and thus,
\begin{equation}
\label{eq:estimuprimeAu}
\NLmu{u'}\, \leq\, 4 \NLmu{{\ensuremath{\mathcal A}} u} + 4K \left( r_0^2+ \pi Z_0^{-1}e^{2Kr_0}(1+8r_0^2) \right)^{\frac 12}\NLmu{u}\, ,
\end{equation}
and by
putting \eqref{eq:estuUmu} and \eqref{eq:estimuprimeAu} together we obtain
\begin{equation}
\begin{split}
\NLmu{u} \, \leq\,& \frac{8\pi}{n-1}\NLmu{{\ensuremath{\mathcal A}} u} + \frac{2\pi}{n-1}4K \left( r_0^2+ \pi Z_0^{-1}e^{2Kr_0}(1+8r_0^2) \right)^{\frac 12}\NLmu{u}\\
&+\frac{n(n+1)}{\pi(n-1)}\Nsqmu{u}\, .
\end{split}
\end{equation}
Let us choose the integer $n=\left\lfloor 16\pi K\left( r_0^2+ \pi Z_0^{-1}e^{2Kr_0}(1+8r_0^2) \right)^{\frac 12} +1\right\rfloor$ so that
\begin{equation}\frac{2\pi}{n-1}4K \left( r_0^2+ \pi Z_0^{-1}e^{2Kr_0}(1+8r_0^2) \right)^{\frac 12}\, \leq\, {\frac 12}\, .\end{equation} In
this
case, we
obtain:
\begin{align}
\NLqmu{u} &\, \leq\, \frac{e^{2Kr_0}}{4K\left( r_0^2+ \pi Z_0^{-1}e^{2Kr_0}(1+8r_0^2) \right)^{\frac 12}}\Nsqmu{Au}\nonumber\\&+
\frac{e^{2Kr_0}\left(16 K\left(r_0^2+ \pi Z_0^{-1}e^{2Kr_0}(1+8r_0^2)\right)^{\frac 12}
+3\right)^2}{16\pi^2 K\left(r_0^2+ \pi Z_0^{-1}e^{2Kr_0}(1+8 r_0^2)\right)^{\frac 12}}\Nsqmu{u}\, ,
\end{align}
which is precisely the inequality \eqref{eq:estim2Bu} we wanted to prove. Inequalities \eqref{eq:estim1Bu} and
\eqref{eq:estim2Bu} give the result, for $a_{K, \delta}:= c^{(1)}_{K, \delta} \cdot c_K^{(3)}$ and $b_{K, \delta}:= c^{(1)}_{K, \delta} \cdot
c_K^{(2)}$.\qedhere
\end{proof}
\begin{proposition}
\label{prop:Lcompactresolvent}
For all $K>1$, there exists $\delta_{3}(K)>0$ such that for all $0<\delta\leq\delta_{3}(K)$, the operator $L^\omega_{{q}}$ is closable. In
that case, its closure has the same domain as the closure of $A$.
\end{proposition}
\begin{proof}
Let us choose $\delta_3(K)>0$ so that
\begin{equation}
\label{eq:condomega37}
b_{K, \delta_3(K)}<1
\end{equation}
where $b_{K, \delta}$ is the constant introduced in \eqref{eq:BAboundedmu}, then, for
all $0<\delta\leq\delta_3(K)$, the operator $B$ is $A$-bounded with $A$-bound strictly lower than $1$. The result is then a
consequence of Th. IV-1.1, p.190 in \cite{Kato1995}.
\end{proof}
\subsubsection{The spectrum of $L^\omega_{{q}}$}
We divide our study into two parts: the determination of the position of the spectrum within a sector and its position near $0$.
\medskip
\subsubsection{Position of the spectrum away from $0$} We prove mainly that the perturbed operator $L^\omega_{{q}}$ still generates an analytic
semigroup of operators on an appropriate sector. An immediate corollary is the fact that the spectrum lies in a cone whose vertex is zero.
We know (Proposition \ref{prop:Apositivemu}) that for all $0<\alpha<\frac{\pi}{2}$, $A$ generates an analytic
semigroup of operators on $\Delta_\alpha:=\ens{\lambda\in{\ensuremath{\mathbf C}} }{|\arg(\lambda)|<\alpha}$.
\begin{proposition}
\label{prop:semigroupAB}
For all $K>1$, $0<\alpha<\frac{\pi}{2}$ and $\gep>0$, there exists $\delta_4>0$ (depending on $\alpha$, $K$ and $\gep$) such that for
all $0<\delta<\delta_{4}$, the spectrum of $L^\omega_{{q}}= A+B$ lies within $\Theta_{\gep,
\alpha}:= \ens{\lambda\in{\ensuremath{\mathbf C}} }{\frac{\pi}{2} + \alpha\leq
\arg(\lambda)\leq \frac{3\pi}{2}-\alpha}\cup \ens{\lambda\in{\ensuremath{\mathbf C}} }{|\lambda|\leq \gep}$. Moreover, there exists $\alpha'\in(0, \frac\pi2)$ such
that the operator $L^\omega_{{q}}$ still generates an analytic semigroup on $\Delta_{\alpha'}$.
\end{proposition}
\begin{proof}[Proof of Proposition \ref{prop:semigroupAB}]
Let $0<\alpha<\frac{\pi}{2}$ be fixed.
Following \eqref{eq:BAboundedmu} and using \eqref{eq:estim R lambda alpha without gep}, one can easily deduce an estimate on the bounded operator
$BR(\lambda, A)$, for $\lambda\in\Sigma_\alpha$:
\begin{equation}
\begin{split}
\Nsqmu{BR(\lambda, A)u} &\,\leq\, a_{K,\delta}\Nsqmu{R(\lambda, A)u} + b_{K,
\delta}\Nsqmu{A R(\lambda,A)u} \\
&\, \leq\, a_{K, \delta}\frac{1}{(1-\sin(\alpha))|\lambda|}\Nsqmu{u}\\ &
\ \ \ \ \ \
+ b_{K, \delta}\left(1+
\frac{1}{1-\sin(\alpha)}\right) \Nsqmu{u}\, .
\end{split}
\end{equation}
Let us fix $\gep>0$ and choose $\delta$ so that:
\begin{equation}
\label{eq:condomega00}
\max\left(4b_{K, \delta}\left(\frac{1}{1-\sin(\alpha)}+1\right) , \frac{4 a_{K, \delta}}{(1-\sin(\alpha))\gep} \right) \,
\leq\, 1\, .
\end{equation}
Then, for $\lambda\in\Sigma_\alpha$ such that $|\lambda|>\gep\geq \frac{4a_{K, \delta}}{1-\sin(\alpha)}$, we have
\begin{equation}\Nsqmu{BR(\lambda, A)u} \, \leq\, \frac{1}{2}\Nsqmu{u}\, .\end{equation}
In particular, $1 - BR(\lambda, A)$ is invertible with $\Nsqmu{\left( 1 - BR(\lambda, A) \right)^{-1}}\leq 2$.
A direct calculation shows that \begin{equation}\left( \lambda - (A+B) \right)^{-1} \, =\, R(\lambda, A) \left( 1 -
BR(\lambda, A)\right)^{-1}\, .\end{equation} One deduces the following estimates on the resolvent: for $\lambda\in\Sigma_\alpha$,
$|\lambda|>\gep$,
\begin{equation}\label{eq:estim resolvent Lq gep prime}\Nsqmu{R(\lambda, L^\omega_{{q}})}\, \leq\, \frac{2}{(1-\sin(\alpha))|\lambda|}\,
.\end{equation}
Estimate \eqref{eq:estim resolvent Lq gep prime} has two consequences: firstly, one deduces immediately that the spectrum $\sigma(L^\omega_{q})$ of
$L^\omega_{q}$ is contained in $\Theta_{\gep,\alpha}$:
\begin{equation}
\label{eq:subset rho Lq}
\sigma(L_q^{\omega})\subseteq\ens{\lambda\in{\ensuremath{\mathbf C}} }{\frac{\pi}{2} + \alpha\leq \arg(\lambda)\leq \frac{3\pi}{2}-\alpha}\cup
\ens{\lambda\in{\ensuremath{\mathbf C}} }{|\lambda|\leq \gep}.
\end{equation}
Secondly, \eqref{eq:estim resolvent Lq gep prime} entails that $L^\omega_{q}$ generates an analytic semigroup of operators on an appropriate
sector. Indeed, if one denotes by $L^{\omega}_{q, \gep}:= L^{\omega}_{q} -\gep$, one deduces from \eqref{eq:subset rho Lq} that
$0\in\rho(L^{\omega}_{q, 2\gep})$ and that for all $\lambda\in{\ensuremath{\mathbf C}} $ with $\Re(\lambda)>0$ (in particular, $|\lambda|<|\lambda+2\gep|$)
\begin{align}
\Nsqmu{R(\lambda, L^\omega_{{q,2\gep}})}\, &=\, \Nsqmu{R(\lambda+2\gep, L^\omega_{{q}})}\, \leq\, \frac{2}{(1-\sin(\alpha))|\lambda+2\gep|}\,
,\nonumber\\
&\, \leq\, \frac{2}{(1-\sin(\alpha))|\lambda|}\, .
\end{align}
Hence, using the same arguments of Taylor expansion as in the proof of Proposition \ref{prop:Apositivemu} and applying Proposition
\ref{prop:pazysemgps}, one easily sees that $L^\omega_{{q,2\gep}}$ generates an analytic semigroup in a (a priori) smaller sector
$\Delta_{\alpha'}$, where $\alpha'\in(0, \frac\pi2)$ can be chosen as $\alpha':= \frac12 \arctan\left( \frac{1-\sin(\alpha)}{2} \right)$. But if
$L^\omega_{{q,2\gep}}$ generates an analytic semigroup, so does $L^\omega_{q}$.
\end{proof}
\subsubsection{Position of the spectrum near $0$}
\label{subsubsec:loczero}
Let us apply Proposition \ref{prop:semigroupAB} for fixed $K>1$, $\alpha\in(0, \frac\pi2)$, $\rho\in(0,1)$ and $\gep:=\rho\gl_{K}$, where we recall
that $\gl_{K}$ is the spectral gap between the eigenvalue $0$ for the non perturbed operator $A$ and the rest of the spectrum
$\sigma(A)\smallsetminus\{0\}$.
Let $\Theta_{\gep, \alpha}^{+}:= \ens{\lambda\in\Theta_{\gep, \alpha}}{\Re(\lambda)\geq 0}$ be the subset of
$\Theta_{\gep, \alpha}$ which lies in the positive part of the complex plane (see Fig. \ref{fig:Thetaeps}). In order to show
the linear stability, one has
to make sure that one can choose a perturbation $B$ small enough so that no eigenvalue of $A+B$ remains in the small set
$\Theta_{\gep, \alpha}^{+}$.
\begin{figure}[!ht]
\centering
\includegraphics[width=0.5\textwidth]{coin.eps}
\caption{The set $\Theta_{\gep,
\alpha}$.}
\label{fig:Thetaeps}
\end{figure}
Since $\gl_{K}>0$, one can separate $0$ from the rest of the spectrum of $A$ by a circle $\mathscr{C}$ centered in $0$ with radius
$(\frac{\rho+1}{2})\gl_{K}$. The appropriate choice of $\gep$ ensures that the
interior of the disk delimited by $\mathscr{C}$ contains $\Theta_{\gep, \alpha}^{+}$ (see Figure~\ref{fig:Thetaeps}).
The main argument is the following: by construction of $\mathscr{C}$, $0$ is the only eigenvalue (with multiplicity $1$) of the
non-perturbed operator $A$ lying in the interior of $\mathscr{C}$. A principle of local continuity of eigenvalues shows that, while
adding a sufficiently small perturbation $B$ to $A$, the interior of $\mathscr{C}$ still contains exactly one eigenvalue (which is \emph{a
priori}
close but not equal to $0$) with the same multiplicity.
But we already know that for the perturbed operator $L^\omega_{{q}}=A+B$, $0$ is always an eigenvalue (since $L^\omega_{{q}}q'=0$). One
can therefore
conclude that, by uniqueness, $0$ is the only element of the spectrum of $L^\omega_{{q}}$ within $\mathscr{C}$, and is an eigenvalue with
multiplicity $1$. In particular, there is no element of the spectrum in the positive part of the complex plane.
In order to quantify the appropriate size of the perturbation $B$, one has to have explicit estimates on the resolvent $R(\lambda,
A)$ on the
circle $\mathscr{C}$.
\begin{lemma}
\label{lem:estimRGamma}
There exists some explicit constant $c_{\mathscr{C}}=c_{\mathscr{C}}(K, \rho)$ such that for all $\lambda\in\mathscr{C}$,
\begin{align}
\Nsqmu{R(\lambda, A)} &\, \leq\, c_{\mathscr{C}}\, ,\label{eq:estimRcercle}\\
\Nsqmu{AR(\lambda, A)} &\, \leq\, 1+ \left(\frac{1+\rho}{2}\right)\gl_{K}\cdot c_{\mathscr{C}}\, .\label{eq:estimARcercle}
\end{align}
One can choose $c_{\mathscr{C}}$ as $\frac{1}{\gl_{K}} \max\left( \frac{2}{\rho+1},\, \frac{2}{1-\rho}\right):= \frac{\ell(\rho)}{\gl_{K}}$.
\end{lemma}
\begin{proof}[Proof of Lemma \ref{lem:estimRGamma}]
Applying the spectral theorem (see \cite[Th. 3, p.1192]{cf:Dunford}) to the essentially self-adjoint operator $A$, there exists a
spectral measure $E$ vanishing on the complementary of the spectrum of $A$ such that $A=\int_{\bbR} \lambda \dd E(\lambda)$. In that
extent, one has for any $\zeta\in\mathscr{C}$
\begin{equation}
R(\zeta, A)\, =\, \int_{\bbR} \frac{\dd E(\lambda)}{\lambda -\zeta}\, .
\end{equation}
In particular, for $\zeta\in\mathscr{C}$
\begin{equation}
\Nsqmu{R(\zeta, A)} \leq \sup_{\lambda\in \sigma(A)} \frac{1}{|\lambda-\zeta|}\leq \frac{\ell(\rho)}{\gl_{K}}\, .
\end{equation}
The estimation \eqref{eq:estimARcercle} is straightforward.
\end{proof}
We are now in position to apply our argument of local continuity of eigenvalues: Following \cite[Th III-6.17,
p.178]{Kato1995}, there exists a decomposition of the
operator $A$ according to $H^{-1}_{{q_0},\mu}= H_0 \oplus H'$ (in
the
sense that $AH_0\subset H_0$, $A H'\subset H'$ and $P {\ensuremath{\mathcal D}} (A) \subset {\ensuremath{\mathcal D}} (A)$,
where $P$ is the projection on $H_0$ along $H'$) in such a way that $A$ restricted to $H_0$ has spectrum
$\{0\}$ and $A$ restricted to $H'$ has spectrum $\sigma(A)\smallsetminus\{0\}$.
Let us note that the dimension of $H_0$ is $1$, since the characteristic space of $A$ in the eigenvalue $0$ is reduced to its
kernel which is of dimension $1$.
Then, applying \cite[Th. IV-3.18, p.214]{Kato1995}, and using Proposition \ref{prop:BAboundedmu},
we find that if one chooses $\delta>0$, such that
\begin{equation}
\label{eq:condperturbGamma}
\sup_{\lambda\in\mathscr{C}} \left(a_{K, \delta} \Nsqmu{R(\lambda, A)} + b_{K, \delta}\Nsqmu{AR(\lambda,
A)}\right)<1,
\end{equation}
then the perturbed operator $L^\omega_{{q}}$ is likewise decomposed according to
$H^{-1}_{{q_0},\mu}= \tilde{H}_0\oplus \tilde{H}'$, in such a way that $\dim(H_0)= \dim(\tilde{H}_0)= 1$, and that the spectrum of
$L^\omega_{{q}}$
is again separated in two parts by $\mathscr{C}$ . But we already know that the characteristic space of the
perturbed operator $L^\omega_{{q}}$ according to the eigenvalue $0$ is, at least, of dimension $1$ (since $L^\omega_{{q}}q'=0$).
We can conclude, that for such an $\delta>0$, $0$ is the only eigenvalue in $\mathscr{C}$ and that $\dim(\tilde{H}_0)=1$.
Applying Lemma \ref{lem:estimRGamma}, we see that condition \eqref{eq:condperturbGamma} is satisfied if we choose $\delta>0$
so that:
\begin{equation}
\label{eq:condomega99}
a_{K, \delta} c_{\mathscr{C}} + b_{K, \delta}\left( 1 + \left( \frac{1+\rho}{2} \right)\gl_{K} c_{\mathscr{C}}\right)<1.
\end{equation}
In particular, in that case, the spectrum of $L^\omega_{{q}}$ is contained in
\begin{equation}
\ens{\lambda\in{\ensuremath{\mathbb C}} }{\frac{\pi}{2} + \alpha\leq
\arg(\lambda)\leq \frac{3\pi}{2}-\alpha}\subseteq
\ens{z\in{\ensuremath{\mathbf C}} }{\Re(z)\leq 0}\, .
\end{equation}
Finally, the following proposition sums-up the sufficient conditions on $\delta$ for the conclusions of Theorem \ref{th:spectral
prop L disorder} to be satisfied:
\begin{proposition}
\label{prop:resume conditions delta_2}
Recall the definitions of $a_{K, \delta}$ and $b_{K, \delta}$ in Proposition \ref{prop:BAboundedmu}.
If $\delta>0$ satisfies the following conditions
\begin{equation}
\label{eq:conds4}
\begin{split}
b_{K, \delta} &\, \leq\, 1\, ,\\
4b_{K, \delta}\left( \frac{1}{1-\sin(\alpha)} +1\right) &\, \leq\, 1\, ,\\
\frac{4 a_{K, \delta} }{\rho\gl_{K}\left( 1-\sin(\alpha) \right)}&\, \leq\, 1\, ,\\
a_{K, \delta}\frac{\ell(\rho)}{\gl_{K}} + b_{K, \delta} \left( 1+ \left(\frac{1+\rho}{2}\right)\ell(\rho)\right)&\, <\, 1\, .
\end{split}
\end{equation}
the conclusions of Theorem \ref{th:spectral prop L disorder} are true.
\end{proposition}
\medskip
\begin{proof}
One has simply to sum-up conditions \eqref{eq:condomega37}, \eqref{eq:condomega00} with $\gep=\rho\gl_{K}$ and
\eqref{eq:condomega99}. \eqref{eq:estim gd K} can be obtained by (long) estimations on the coefficients $a_{K, \delta}$ and $b_{K, \delta}$.
\end{proof}
\medskip
\begin{rem}\rm
The conditions in Proposition~\ref{prop:resume conditions delta_2}
can be simplified. For example one can exhibit an explicit
constant $c$ such that if $\delta$
satisfies
\begin{equation}
\label{eq:estim gd K}
\begin{split}
\delta e^{12\pi\delta}\,\leq\, c e^{-20K\bar{r_\delta}}\max&\left(1, \left( \frac{1-\sin(\alpha)}{2-\sin(\alpha)} \right),
\frac{\rho\gl_{K}(1-\sin(\alpha))e^{-4K\bar{r_\delta}}}{K^2},\right.\\
&\left.\ \ \ \ \frac{\gl_{K}}{K^2e^{4K\bar{r_\delta}}\ell(\rho) + \gl_{K} \left( 1+ \left( \frac{1+\rho}{2} \right)\ell(\rho) \right)}\right)
\end{split}
\end{equation}
the conditions in \eqref{eq:conds4} are fulfilled. Explicit estimates on the spectral gap $\lambda_K$
can be found
in \cite[Sec.~2.5]{cf:BGP}.
\end{rem}
\section*{Acknowledgments}
We are grateful to K. Pakdaman and G. Wainrib for helpful discussions.
G. G. acknowledges the support of the ANR (projects Mandy and SHEPI) and
the support of the Petronio Fellowship Fund at the Institute for Advanced Study
(Princeton, NJ) where this work has been completed.
\begin{appendix}
\section{Regularity in the non-linear Fokker-Planck equation}
\label{sec:appendix regularity pt with disorder}
The purpose of this section is to establish regularity properties of the solution of the non-linear
equation \eqref{eq:AR} (where we fix $\delta=1$ for simplicity). Note that this case also captures the situation where $U(\cdot, \omega)\equiv
\omega$ (evolution \eqref{FKP kuramoto disorder}), as well as the situation where $U(\cdot, \cdot)\equiv 0$ (evolution \eqref{FKP kuramoto disorder
without drift}). In what follows we make the assumption that $U$ is bounded and that for all $\omega\in\Supp(\mu)$, $\theta\mapsto
U(\theta, \omega)\in C^\infty({\ensuremath{\mathbb S}} ; \, {\ensuremath{\mathbb R}} )$ with bounded derivatives.
The existence and uniqueness in $L^2(\lambda\otimes\omega)$ of a solution to \eqref{eq:AR} can be tackled using Banach fixed point arguments
(see \cite[Section 4.7]{cf:SellYou}), but one can obtain more regularity from the theory of fundamental solutions of parabolic equations.
More precisely, it is usual to interpret Equation \eqref{eq:AR} as the strong formulation of the weak equation (where $\nu \in{\ensuremath{\mathcal C}} ([0, T],
{\ensuremath{\mathcal M}} _{1}({\ensuremath{\mathbb S}} \times{\ensuremath{\mathbb R}} ))$ and $F$ is any bounded function on ${\ensuremath{\mathbb S}} \times{\ensuremath{\mathbb R}} $ with twice bounded derivatives w.r.t. $\theta$):
\begin{align}
\int_{\bbR\times{\bbS }} F(\theta, \omega)\nu_t(\dd\theta, \dd\omega)&= \int_{\bbR\times{\bbS }} F(\theta, \omega)\nu_0(\dd\theta, \dd\omega) +\frac12\int_0^t \int_{\bbR\times{\bbS }} F''(\theta, \omega)\nu_s(\dd\theta,
\dd\omega)\dd s\nonumber\\
&+ \int_0^t \int_{\bbR\times{\bbS }} F'(\theta, \omega) \left(\int_{\bbR\times{\bbS }} J(\theta-\cdot)\dd\nu_s+ U(\theta, \omega)\right)\nu_s(\dd\theta,\dd\omega)\dd s,
\label{eq:weak formulation qt with U}
\end{align}
where the second marginal (w.r.t. to the disorder $\omega$) of the initial condition $\nu_0(\dd\theta, \dd\omega)$ is $\mu(\dd\omega)$ so that one can
write
\begin{equation}
\label{eq:initial condition nu zero}
\nu_0(\dd\theta, \dd\omega) = \nu_0^\omega(\dd\theta) \mu(\dd\omega)\, ,
\end{equation}
where $\nu_0^\omega$ is a probability measure on ${\ensuremath{\mathbb S}} $, for $\mu$-a.e. $\omega$.
As already mentioned, a proof of the existence of a solution on $[0, T]$ of \eqref{eq:weak formulation qt with U} can be obtained
from the almost-sure convergence of the empirical measure of the microscopic system \cite{cf:eric}. One can also find a proof of uniqueness of
such a solution relying on arguments introduced in \cite{cf:Oelschlager}.
The regularity result can be stated as follows:
\begin{proposition}
\label{prop:regularity solution qt}
For all probability measure $\nu_0(\dd\theta, \dd\omega)=\nu_0^\omega(\dd\theta)\mu(\dd\omega)$ on ${\ensuremath{\mathbb S}} \times{\ensuremath{\mathbb R}} $, for all $T>0$, there exists a unique
solution $\nu$ to \eqref{eq:weak formulation qt with U} in
${\ensuremath{\mathcal C}} ([0, T], {\ensuremath{\mathcal M}} _1({\ensuremath{\mathbb S}} \times{\ensuremath{\mathbb R}} ))$ such that for all $F\in{\ensuremath{\mathcal C}} ({\ensuremath{\mathbb S}} \times{\ensuremath{\mathbb R}} )$,
\begin{equation}
\lim_{t\searrow0}\int_{\bbR\times{\bbS }} F(\theta, \omega)\nu_t(\dd\theta, \dd\omega)= \int_{\bbR\times{\bbS }} F(\theta, \omega) \nu_0^\omega(\dd\theta) \mu(\dd\omega).
\end{equation}
Moreover, for all $t>0$, $\nu_t$ is absolutely continuous with respect to $\lambda_1\otimes\mu$ and for $\mu$-a.e. $\omega\in Supp(\mu)$, its
density $(t,\theta, \omega)\mapsto p_t(\theta, \omega)$ is strictly positive on $(0, T]\times {\ensuremath{\mathbb S}} $, is ${\ensuremath{\mathcal C}} ^\infty$ in $(t,
\theta)$ and solves the Fokker-Planck equation \eqref{eq:AR}.\end{proposition}
\begin{proof}[Proof of Proposition \ref{prop:regularity solution qt}]
Let us fix $T>0$, $\omega\in\Supp(\mu)$ and $t\mapsto \nu_t$ the unique solution in ${\ensuremath{\mathcal C}} ([0, T], {\ensuremath{\mathcal M}} _1({\ensuremath{\mathbb S}} \times{\ensuremath{\mathbb R}} ))$ to \eqref{eq:weak
formulation qt with U}. Let us define $R(t, \theta, \omega):= \int_{\bbR\times{\bbS }} J(\theta-\cdot)\dd\nu_t+ U(\theta, \omega)$ and consider the linear equation
\begin{equation}
\label{eq:linear equation qt with Rt}
\partial_t p_t(\theta,\omega)\, =\, \frac{1}{2} \Delta p_t(\theta,\omega) -\partial_\theta \Big(p_t(\theta,\omega)R(t, \theta, \omega)\Big)\, ,
\end{equation}
such that for $\mu$-a.e. $\omega$, for all $F\in{\ensuremath{\mathcal C}} ({\ensuremath{\mathbb S}} )$,
\begin{equation}
\label{eq:linear equation qt with Rt cond limit}
\int_{\bbS} F(\theta)p_t(\theta, \omega) \dd\theta\,
\stackrel{t\searrow 0}{\longrightarrow}
\int_{\bbS} F(\theta) \nu_0^\omega(\dd\theta)\, .
\end{equation}
For fixed $\omega\in\Supp(\mu)$, $R(\cdot, \cdot, \omega)$ is continuous in time and ${\ensuremath{\mathcal C}} ^\infty$ in $\theta$.
Suppose for a moment that we have found a weak solution $p_t(\theta,\omega)$ to \eqref{eq:linear equation qt with Rt}-\eqref{eq:linear equation qt
with Rt cond limit} such that for
$\mu$-a.e. $\omega$, $p_t(\cdot, \omega)$ is strictly positive on $(0,T]\times{\ensuremath{\mathbb S}} $. In particular for such a solution $p$, the quantity
$\int_{\bbS} p_t(\theta, \omega)\dd\theta$ is conserved for $t>0$, so that $p_t(\cdot, \omega)$ is indeed a probability density for all $t>0$. Then
both
probability measures $\nu_t(\dd\theta, \dd\omega)$ and
$p_t(\theta,\omega)\dd\theta\mu(\dd\omega)$ solve
\begin{align}
\int_{\bbR\times{\bbS }} F(\theta, \omega)\nu_t(\dd\theta, \dd\omega)&= \int_{\bbR\times{\bbS }} F(\theta, \omega)\nu_0(\dd\theta, \dd\omega) +\frac12\int_0^t \int_{\bbR\times{\bbS }} F''(\theta, \omega)\nu_s(\dd\theta,
\dd\omega)\dd s\nonumber\\
&+ \int_0^t \int_{\bbR\times{\bbS }} F'(\theta, \omega) R(t, \theta, \omega)\nu_s(\dd\theta,\dd\omega)\dd s.
\label{eq:weak formulation qt with Rt}
\end{align}
By \cite{cf:eric} or \cite[Lemma 10]{cf:Oelschlager}, uniqueness in \eqref{eq:weak formulation qt with U} is precisely a consequence
of uniqueness in \eqref{eq:weak formulation qt with Rt}. Hence, by uniqueness in \eqref{eq:weak formulation qt with Rt},
$\nu_t(\dd\theta, \dd\omega)=p_t(\theta,\omega)\dd\theta\mu(\dd\omega)$, which is the result. So it suffices to exhibit a weak solution
$p_t(\theta,\omega)$ to \eqref{eq:linear equation qt with Rt} such that \eqref{eq:linear equation qt with Rt cond limit} is satisfied.
\medskip
This fact can be deduced from standard results for uniform parabolic PDEs (see \cite{cf:Aronson} and \cite{cf:Friedman} for precise definitions).
In particular, a usual result, which can be found in \cite[\S 7 p.658]{cf:Aronson}, states that \eqref{eq:linear equation qt with Rt} admits a
fundamental solution $\Gamma(\theta, t; \theta', s, \omega)$ ($t>s$),
which is bounded above and below (see \cite[Th.7, p.661]{cf:Aronson}):
\begin{equation}
\label{eq:control Gamma}
\frac{1}{C\sqrt{t-s}} \exp\left( \frac{-C(\theta - \theta')^2}{\sqrt{t-s}}\right)\leq \Gamma(\theta, t; \theta', s, \omega) \leq \frac{C}{\sqrt{t-s}}
\exp\left( \frac{-(\theta - \theta')^2}{C\sqrt{t-s}}\right)\, .
\end{equation}
Note that the constant $C>0$ only depends on $T$ and the \emph{structure} of the linear operator in \eqref{eq:linear equation qt with
Rt} (see \cite[Th.7, p.661]{cf:Aronson} and \cite[\S 1, p.615]{cf:Aronson}). In particular, since $(\theta, \omega)\mapsto U(\theta, \omega)$ is
bounded, this constant does not depend on $\omega$.
Note that the proof given in \cite{cf:Aronson} is done for $\theta\in{\ensuremath{\mathbb R}} $ but can be readily adapted to our case ($\theta\in{\ensuremath{\mathbb S}} $).
Moreover, thanks to Corollary 12.1, p.690 in \cite{cf:Aronson}, the following expression of $p_t(\theta, \omega)$
\begin{equation}
\label{eq:pt Gamma}
p_t(\theta, \omega)=\int_{\bbS} \Gamma(\theta, t; \theta', 0, \omega)\nu_0^\omega(\dd\theta')
\end{equation}
defines a weak solution of \eqref{eq:linear equation qt with Rt} on $(0,T]\times{\ensuremath{\mathbb S}} $ (namely a weak solution on $(\tau, T]\times {\ensuremath{\mathbb S}} $, for all
$0<\tau<T$) such that \eqref{eq:linear equation qt with Rt cond limit} is satisfied. The positivity and boundedness of $p_t(\cdot, \omega)$ for
$t>0$ is an easy consequence of \eqref{eq:control Gamma}. The smoothness of $p_\cdot(\cdot, \omega)$ on $(0, T]\times {\ensuremath{\mathbb S}} $ can be derived by
standard bootstrap methods.
\end{proof}
We focus now on the regularity of the solution $p_t(\theta, \omega)$ of \eqref{eq:AR} with respect to the disorder $\omega$. We assume here that the
initial condition $\nu_0$ is such that for all $\omega\in\Supp(\mu)$, $\nu_0^\omega(\dd\theta)$ is absolutely continuous with respect to the Lebesgue
measure $\lambda_1$ on ${\ensuremath{\mathbb S}} $: there exists a positive integrable function $\gamma(\cdot, \omega)$ of integral $1$ on ${\ensuremath{\mathbb S}} $ such that
$\nu_0^\omega(\dd\theta)=\gamma(\theta, \omega)\dd\theta$. Then we have
\begin{lemma}[Regularity w.r.t. the disorder]
\label{lem:regularity density disorder}
For every $(t_0, \theta_0)\in(0, \infty)\times{\ensuremath{\mathbb S}} $, for every $\omega_0$ which is an accumulation point in $\Supp(\mu)$ such that the following
holds
\begin{equation}
\label{eq:condition continuity gamma}
\int_{\bbS} |\gamma(\theta, \omega)-\gamma(\theta, \omega_0)|\dd\theta \to 0, \quad\text{as $\omega\to\omega_0$}\, ,
\end{equation}
then the solution $p$ of \eqref{eq:AR} defined on $(0, \infty)\times{\ensuremath{\mathbb S}} \times \Supp(\mu)$ is continuous at the point $(t_0, \theta_0, \omega_0)$.
\end{lemma}
\begin{proof}[Proof of Lemma \ref{lem:regularity density disorder}]
For any $\omega$ in the support of $\mu$, let for all $t>0$, $\theta\in{\ensuremath{\mathbb S}} $
\begin{equation}
\label{eq:def u diff q1 q2}
u(t, \theta, \omega) := p_t(\theta, \omega) - p_t(\theta, \omega_0),
\end{equation}
where $(p_t(\cdot, \cdot))_{t\geq 0}$ is the unique solution of \eqref{eq:linear equation qt with Rt}. It is easy to see that $u$ is a strong
solution to the following PDE
\begin{equation}
\label{eq:pde verified by u}
\partial_t u(t, \theta, \omega) - \left[\frac12 \Delta u(t, \theta) - \partial_\theta\left( u(t, \theta) R(t, \theta, \omega_0)\right)\right] =
\mathcal{R}(t, \theta, \omega),
\end{equation}
where $\mathcal{R}(t, \theta, \omega):= \partial_\theta\left[ p_t(\theta, \omega)\left( R(t, \theta, \omega) -R(t, \theta, \omega_0) \right)\right]$ and
with initial condition (since $\nu_0^\omega(\dd\theta)=\gamma(\theta, \omega)\dd\theta$ for all $\omega$)
\begin{equation}
\label{eq:initial condition u}
u(t, \theta, \omega)|_{t\searrow0} =\gamma(\theta, \omega) - \gamma(\theta, \omega_0).
\end{equation}
Then applying \cite[Th. 12 p.25]{cf:Friedman}, $u(t, \theta, \omega)$ can be expressed as
\begin{equation}
\label{eq:u Gamma}
u(t, \theta, \omega)=\int_{\bbS} \Gamma(\theta, t; \theta', 0, \omega_0)(\gamma(\theta, \omega) - \gamma(\theta, \omega_0))\dd\theta' - \int_0^t \int_{\bbS} \Gamma(\theta,
t;\theta', s, \omega_0)\mathcal{R}(s, \theta', \omega)\dd\theta'\dd s.
\end{equation}
For the first term of the RHS of \eqref{eq:u Gamma}, we have
\begin{equation}
\left|\int_{\bbS} \Gamma(\theta, t; \theta', 0, \omega_0)(\gamma(\theta, \omega) - \gamma(\theta, \omega_0))\dd\theta'\right|\leq
\frac{C}{\sqrt{t}}\int_{\bbS}|\gamma(\theta, \omega) - \gamma(\theta, \omega_0)|\dd\theta'\, ,
\end{equation}
which converges to $0$, for fixed $t>0$, by hypothesis \eqref{eq:initial condition u}.
Secondly, it is easy to see from the definition \eqref{eq:pt Gamma} of the density $p$ and the estimates \eqref{eq:control Gamma} and \cite[Th.9
p.263]{cf:Friedman} concerning the fundamental solution $\Gamma$ that both $p_t(\theta, \omega)$ and $\partial_\theta
p_t(\theta, \omega)$ are bounded uniformly on $(t, \theta, \omega)\in [0, T]\times {\ensuremath{\mathbb S}} \times \Supp(\mu)$. In particular, a
standard result shows that for fixed $(t, \theta)$, the second term of the RHS of \eqref{eq:u Gamma} goes to $0$ as
$\omega\to\omega_0$. But then the joint continuity of $p$ at $(t_0,\theta_0, \omega_0)$ follows from \eqref{eq:pt Gamma} and uniform estimates on $\Gamma$
(see \cite[Th.9 p.263]{cf:Friedman}).
\end{proof}
\end{appendix}
|
\section{Introduction}
We establish lower bounds for the inner and
outer errors in internal diffusion limited aggregation (internal DLA).
Internal DLA models a discrete cluster growth, and is defined as follows.
Let $\Lambda$ be a subset of ${\mathbb Z}^d$ which represents
the explored region at time 0. Let $N$ be an integer,
and $\xi=(\xi_1,\dots, \xi_N)$ be the initial positions
of $N$ independent simple random walks $S_1,\dots,S_N$ on ${\mathbb Z}^d$.
The cluster of volume $N$, denoted $A(\Lambda, \xi)$
is obtained inductively as follows. First, $A(\Lambda,0)=\Lambda$.
Now, assume $A(\Lambda,k-1)$ is obtained, and define
\be{time-settling}
\tau_k=\inf\acc{t\ge 0:\ S_k(t)\not\in A(\Lambda,k-1)},\quad\text{and}\quad
A(\Lambda,k)=A(\Lambda,k-1)\cup \{S_k(\tau_k)\}.
\end{equation}
We call explorers the random walks obeying the aggregation rule
\reff{time-settling}. We say that the explorer $k$ settles
at time $\tau_k$. When $\Lambda=\emptyset$, we also denote
$A(\emptyset,\xi)$ by $A(\xi)$.
In dimensions two and more,
Lawler, Bramson and Griffeath~\cite{lawler92} prove that in order
to cover, without holes, a sphere of radius $n$, we need about the
number of sites of ${\mathbb Z}^d$ in this sphere. In other words,
the asymptotic shape of the cluster
is a sphere. Then, Lawler in~\cite{lawler95} shows a
subdiffusive upper bound for the worse fluctuation to the
spherical shape. More precisely,
the latter result is formulated in terms of inner and outer errors,
which we now introduce with some notation.
We denote with $\|\cdot\|$ the euclidean norm on ${\mathbb R}^d$.
For any $x$ in $R^d$ and $r$ in ${\mathbb R}$, set
\be{ball-dfn}
B(x,r) = \left\{ y\in{\mathbb R}^d :\: \|y-x\| < r \right\}
\quad\mbox{and}\quad {\mathbb B}(x,r) = B(x,r) \cap {\mathbb Z}^d.
\end{equation}
For $\Lambda\subset{\mathbb Z}^d$,
$|\Lambda|$ denotes the number of sites in $\Lambda$.
The inner error $\delta_I(n)$ is such that
\be{def-inner}
n-\delta_I(n)=\sup\acc{r \geq 0:\: {\mathbb B}(0,r)\subset A\big(|{\mathbb B}(0,n)|\big)}.
\end{equation}
Also, the outer error $\delta_O(n)$ is such that
\be{def-outer}
n+\delta_O(n)=\inf\acc{r \geq 0:\: A\big(|{\mathbb B}(0,n)|\big)\subset {\mathbb B}(0,r)}.
\end{equation}
Note that $\delta_I$ and $\delta_O$ represent the worse fluctuations
to the spherical shape.
In 2010, Asselah-Gaudilli\`ere in \cite{AG1,AG2},
and Jerison-Levine-Sheffield in \cite{JLS1,JLS2},
independently showed the following upper bound.
\bt{theo-upper}
When dimension $d\ge 3$,
there are constants $\{\beta_d,\ d\ge 3\}$
such that with probability~1,
\be{theo-res2}
\limsup_{n\to\infty} \frac{\delta_I(n)}{\sqrt{\log(n)}}\le \beta_d,
\quad\text{and}\quad
\limsup_{n\to\infty} \frac{\delta_O(n)}{\sqrt{\log(n)}}\le \beta_d.
\end{equation}
When dimension is $2$,
there is a constant $\{\beta_2\}$ such that with probability~1,
\be{theo-res1}
\limsup_{n\to\infty} \frac{\delta_I(n)}{\log(n)}\le \beta_2,
\quad\text{and}\quad
\limsup_{n\to\infty} \frac{\delta_O(n)}{\log(n)}\le \beta_2.
\end{equation}
\end{theorem}
All these studies are concerned with upper bounds on
$\delta_I$ and $\delta_O$, and the matter of showing that
they were indeed realized remained untouched.
We present now two results on lower bounds. The first one is
independent from previous results.
This bound which characterizes the worse fluctuations
is optimal in $d\ge 3$. Morally, it says that if there are no
{\it deep} holes, then {\it long} tentacles form.
\bp{prop-simple}
When dimension is three or more, set $h(k)= \sqrt{\log(k)}$
there is $\alpha_d>0$ such that
\be{prop-d3}
\lim_{n\to\infty} P\pare{ \exists k>n,\ \delta_O(k)\ge
\alpha_d h(k)\quad\big|\quad
\delta_I(n)< \alpha_dh(n)}=1.
\end{equation}
When dimension is two, set $h(k)= \sqrt{\log(k)\log(\log(k))}$.
There is $\alpha_2>0$ such that
\be{prop-d2}
\lim_{n\to\infty} P\pare{ \exists k>n,\ \delta_O(k)\ge
\alpha_2 h(k)\quad\big|\quad
\delta_I(n)< \alpha_2h(n)}=1.
\end{equation}
\end{proposition}
\br{rem-io}
Note that as a consequence of \reff{prop-d3}, we have
$\alpha_d$ positive such that
\be{main-io}
P\pare{ \delta_O(n)\ge \alpha_d\sqrt{\log(n)} \text{ or }
\delta_I(n)\ge \alpha_d\sqrt{\log(n)},\ \text{ i.o.}}=1,
\end{equation}
\end{remark}
The second result uses the
upper bound of Theorem~\ref{theo-upper}. It says
that {\it deep} holes do form.
\bt{theo-lower}
There are positive constants $\{\alpha_d,d\ge 2\}$ such that
when $d\ge 3$,
\be{main-A}
P\big(\delta_I(n)> \sqrt{ \alpha_d \log(n)}\quad i.o.\big)=1
\end{equation}
\end{theorem}
\br{rem-conj} We believe that in dimension 2,
for some $\alpha_2>0$, almost surely both $\delta_I(n)$ and
$\delta_O(n)$ are larger than $\alpha_2 \log(n)$ infinitely often.
However, the way of realizing this event
is probably different from what
we describe in $d\ge 3$.
\end{remark}
We present now a side result dealing with
fluctuation in a given direction.
The results on fluctuations for internal DLA \cite{lawler92,lawler95,
AG1,AG2,JLS1,JLS2} all focus on inner and outer errors.
The paper \cite{JLS3} if still of a different nature:
it addresses averaged error, in a sense
where inner and outer errors can cancel each other out.
Though interesting,
fluctuations in a given direction remained untreated.
This is presented
as an open problem in Section 3 of \cite{JLS2}, and we note that
the proof of \cite{AG2} yields the following
bound for the inner error in a given direction.
\bp{prop-AG}
There are positive constants $\{\kappa_d,d\ge 2\}$ such
that for $z$ be such that $\|z\|<n$, we have in ${\mathbb Z}^d$
\be{AG-directional}
P\pare{ z\not\in A(N_n)}\le
\left\{ \begin{array}{ll}
\exp\pare{-\kappa_2 \frac{(n-\|z\|)^2}{\log(n)}}& \mbox{ for } d=2 \, , \\
\exp\pare{-\kappa_d (n-\|z\|)^2}& \mbox{ for } d\ge 3 \, .
\end{array} \right.
\end{equation}
\end{proposition}
The rest of the paper is organized as follows.
In Section~\ref{sec-notation}, we set notation and recall
useful results. In Section~\ref{sec-outer}, we deal with
the outer error and prove Proposition
\ref{prop-simple}. In Section~\ref{sec-inner}, we deal with
the inner error, and prove Theorem~\ref{theo-lower}. Finally,
in Section~\ref{sec-last}, we explain how to read
Proposition~\ref{prop-AG} as a corollary of our previous work
\cite{AG2}.
\section{Notation and Prerequisite}\label{sec-notation}
Let $S:{\mathbb N}\to{\mathbb Z}^d$ denotes a simple random
walk on ${\mathbb Z}^d$. When the initial condition
is $S(0)=z\in {\mathbb Z}^d$, its law is denoted ${\mathbb P}_z$.
The first time $S$ hits a domain $\Lambda\subset {\mathbb Z}^d$,
is denoted $H(\Lambda)$, or $H(S;\Lambda)$ to emphasize
that $S$ is the walk.
For a positive $\gamma$,
we denote by $\rho(\gamma)$ the radius of the largest
ball centered at 0 whose volume is less than $\gamma$. In other words,
\be{not1}
\rho(\gamma)=\sup\acc{n\ge 0: |{\mathbb B}(0,n)|\le \gamma}.
\end{equation}
The abbreviation $b(n)=|{\mathbb B}(0,n)|$ is also handy.
We now consider
a configuration $\eta\in {\mathbb N}^{{\mathbb Z}^d}$ with a finite number
of particles
\be{not3}
|\eta|:=\sum_{z\in {\mathbb Z}^d} \eta(z)<\infty.
\end{equation}
For $R>0$, $\Lambda\subset {\mathbb Z}^d$, and
$z\in {\mathbb B}(0,R)$ (resp. $z\in \partial {\mathbb B}(0,R)$),
we denote by
$W_R(\Lambda;\eta,z)$ the number of explorers visiting
$z$ {\it before} they exit ${\mathbb B}(0,R)$ (resp. {\it at the moment}
they reach $\partial {\mathbb B}(0,R)$), when
the explorers start on $\eta$
with an explored region $\Lambda$.
The positions of explorers settled before they exit
${\mathbb B}(0,R)$ is denoted $A_R(\Lambda;\eta)$, which is also the
support of $W_R(\Lambda;\eta,z)$ in ${\mathbb B}(0,R)$.
Note that $A_R(\Lambda;\eta)$ is contained in ${\mathbb B}(0,R)$.
When we build the
internal DLA cluster from $n$ independent random walks,
we also consider the unrestricted trajectories. This gives
a natural coupling between explorers and independent random walks.
Note that if ${\mathbb B}(0,R)\subset \Lambda$ and $z\in \partial {\mathbb B}(0,R)$,
then
\[
W_R(\Lambda;\eta,z)=W_R({\mathbb B}(0,R);\eta,z),
\]
and this corresponds
to the number of independent random walks which exit
${\mathbb B}(0,R)$ at site $z$. For simplicity, we call this number
$M_R(\eta,z)$. In general, under a natural coupling,
for $z\in {\mathbb B}(0,R)\cup\partial {\mathbb B}(0,R)$,
\be{lawler-3}
W_R(\Lambda, \eta,z) \le M_R(\eta,z).
\end{equation}
\subsection{The abelian property}\label{sec-abelian}
Diaconis and Fulton~\cite{diaconis-fulton}
allow explorers to start on distinct sites, and
show that the law of the cluster is invariant under permutation of
the order in which explorers are launched. This invariance is
named {\it the abelian property}. As a consequence,
one can realize the cluster by sending many {\it exploration waves}.
Let us illustrate this observation by building
$A(\emptyset,(n+m)\delta_0)$ in three waves, since
we need later this very example. The first wave
consists in launching $n$ explorers,
and they settle in $A_1=A(\emptyset,n\delta_0)$, which we call
the cluster {\it after the first wave}. Then, we launch $m$
explorers that we color {\it green} for simplicity,
and if they reach $\partial {\mathbb B}(0,R)$ before settling,
then we stop them on $\partial {\mathbb B}(0,R)$.
The settled green explorers make up the cluster $A_R(A_1; m\delta_0)$.
The cluster after the second wave is then
\[
A_2=A_1\cup A_R(A_1; m\delta_0).
\]
For $z\in \partial {\mathbb B}(0,R)$, we call $\zeta_R$
the configuration of the green explorers
stopped on $\partial {\mathbb B}(0,R)$. In other words,
\[
\zeta_R=\acc{W_R(A_1;m\delta_0,z),\ z\in \partial {\mathbb B}(0,R)}.
\]
Then, the cluster after the third wave is obtain as we launch
the stopped green explorers, and
\[
A_3=A_2\cup A(A_2;\zeta_R).
\]
The abelian properties implies that $A(\emptyset,(n+m)\delta_0)$
equals in law to $A_3$. It is convenient to think of the growing cluster
as evolving in discrete time,
where time counts the number of exploration waves.
\subsection{On the harmonic measure.}\label{sec-harmonic}
We first recall a well known property of Poisson variables.
\bl{lem-obvious}
Let $\{U_n,n\in {\mathbb N}\}$ be an i.\,i.\,d.\, sequence with value
in a set $E$, and $\{E_1,\dots,E_n\}$ a partition of $E$.
Then, if $X$ is an independent Poisson random variable
of parameter $\lambda$, and if
\[
X_i=\sum_{n\le X} \hbox{ 1\hskip -3pt I}_{U_n\in E_i},
\]
then $\{X_i,\ i=1,\dots,n\}$ are independent Poisson
variables with $E[X_i]=\lambda\times P(U\in E_i)$.
\end{lemma}
Now, for $z\in {\mathbb Z}^d\backslash\{0\}$, let
\be{not2}
\Sigma(z)=\partial {\mathbb B}(0,\|z\|),
\end{equation}
and note that $z$ belongs to $\Sigma(z)$ since there
is $z'\in {\mathbb B}(0,\|z\|)$ and $\|z-z'\|=1$
(see Lemma 2.1 of \cite{AG1}).
When we start independent random walks on $\eta$, and
for $z\not= 0$, $h>0$, and $\Lambda\subset \Sigma(z)$, we denote
by $N_z(\eta,\Lambda,h)$ the number of walks initially on $\eta$
which hit $\Sigma(z)$ on $\Lambda$, and then visit $z$ before
exiting ${\mathbb B}(0,\|z\|+h)$.
As a consequence of Lemma~\ref{lem-obvious}, we have
\bc{cor-obvious}
Let $X$ be a Poisson variable of parameter $\lambda$, and
$\eta=X\delta_0$. For any $z\not= 0$, $h,h'>0$, and $\Lambda,\Lambda'
\in \Sigma(z)$ with $\Lambda\cap\Lambda'=\emptyset$, we have
that $N_z(\eta,\Lambda,h)$ and $N_z(\eta,\Lambda',h')$ are independent
Poisson variable, and
\be{mean-N}
\begin{split}
E\cro{N_z(\eta,\Lambda,h)}=
&\lambda\times {\mathbb P}_0\Big(S\big(H(\Sigma(z))\big)\in \Lambda,
\ H(z)<H\big({\mathbb B}^c(0,\|z\|+h)\big)\Big)\\
=&\lambda\sum_{y\in \Lambda}{\mathbb P}_0\Big(S\big(H(\Sigma(z))=y\big)\Big)
{\mathbb P}_y\Big(\ H(z)<H\big({\mathbb B}^c(0,\|z\|+h)\big)\Big).
\end{split}
\end{equation}
\end{corollary}
By combining well known asymptotics of the harmonic measure
with Corollary~\ref{cor-obvious}, we obtain the following lemma.
\bl{lem-main}
Assume that dimension is two or more.
Let $\eta$ be as in Corollary~\ref{cor-obvious}.
For $z\not= 0$, and $R>0$ let $\Lambda={\mathbb B}(z,R)\cap \Sigma(z)$, and
$\Lambda'=\Sigma(z)\backslash \Lambda$. There is $\kappa>0$, independent
of $z$, and $R$ such that
\be{not4}
P\big(N_z(\eta,\Lambda,\infty)=0,\ N_z(\eta,\Lambda',R)=0\big)\le
\exp\pare{-\kappa \frac{\lambda R}{\|z\|^{d-1}}}.
\end{equation}
\end{lemma}
\begin{proof}
Since $N_z(\eta,\Lambda,\infty)$ and $N_z(\eta,\Lambda',R)$
are independent Poisson variables, we have
\be{not5}
P\big(N_z(\eta,\Lambda,\infty)=0,\ N_z(\eta,\Lambda',R)=0\big)=
\exp\big(-E\cro{N_z(\eta,\Lambda,\infty)}-E\cro{
N_z(\eta,\Lambda',R)}\big).
\end{equation}
It remains to compute expected values.
To estimate $E\cro{N_z(\eta,\Lambda,\infty)}$, we recall that
there is a constant $c_d$ such that
\be{not6}
{\mathbb P}_y(H(z)<\infty)\le \frac{c_d}{1+\|y-z\|^{d-2}}.
\end{equation}
Using \reff{mean-N}, there is a constant $c$
\be{not7}
\begin{split}
E\cro{N_z(\eta,\Lambda,\infty)}\le &
\sum_{y\in \Lambda} \lambda{\mathbb P}_0\Big(S\big(H(\Sigma(z))\big)=y\Big)\times
\frac{c_d}{1+\|y-z\|^{d-2}}\\
\le & \frac{c}{2}\frac{\lambda}{\|z\|^{d-1}}
\Big(1+\sum_{k=1}^{R} \frac{k^{d-2}}{1+k^{d-2}}\Big)\le c
\frac{\lambda R}{\|z\|^{d-1}}.
\end{split}
\end{equation}
Now, to estimate $E\cro{N_z(\eta,\Lambda',R)}$, we recall
Lemma 5(b) of \cite{JLS1}, which states that for some
constant $c_d'$, for $y\in \Sigma(z)$
\be{not8}
{\mathbb P}_y\pare{ H(z)< H\big( {\mathbb B}^c(0,\|z\|+R)\big)}\le
\frac{ c_d' R^2}{\|z-y\|^d}
\end{equation}
Using \reff{mean-N} and \reff{not8}, there is a constant $c'$ such that
\be{not9}
\begin{split}
E\cro{N_z(\eta,\Lambda',R)}\le &\sum_{y\in \Sigma(z)\backslash {\mathbb B}(z,R)}
\lambda {\mathbb P}_0\Big(S\big(H(\Sigma(z))\big)=y\Big)
\frac{ c_d' R^2}{\|z-y\|^d}\\
\le &\ \frac{c'}{2} \frac{\lambda R^2}{\|z\|^{d-1}}
\sum_{k=R}^{2\|z\|+2} \frac{k^{d-2}}{k^d}\le \
c' \frac{\lambda R}{\|z\|^{d-1}}
\end{split}
\end{equation}
Combining \reff{not7}, and \reff{not9}, we obtain the desired result.
\end{proof}
\section{The outer error}\label{sec-outer}
In this section, we prove Proposition~\ref{prop-simple}.
Let us explain the proof
in dimension three or more, and explain in Remark~\ref{rem-d2}
of Step 2 how we adapt the proof to dimension two.
For positive reals $\alpha$ and $\gamma$, to be chosen later,
we set $h(n)=\alpha\sqrt{\log(n)}$ and $\sous{L}(n)=
\gamma\sqrt{\log(n)}$ for estimates on the outer and inner
fluctuations.
Even though we eventually take $h(n)=\sous{L}(n)$,
it is useful to keep in mind their distinct nature.
The limit \reff{main-io} follows if for some
small $\gamma=\alpha$, we have
\be{main-1}
\lim_{n\to\infty} P\pare{ \exists k\ge n,\ \delta_O(k)\ge h(k)
\ \big|\ \delta_I(n)<\sous{L}(n)}=1.
\end{equation}
Indeed,
\be{main-2}
\begin{split}
P\big( \exists k\ge n,&\ \delta_O(k)\ge h(k)\quad \text{ or }\quad
\delta_I(k)\ge \sous{L}(k)\big)\\
\ge& P\big(\delta_I(n)\ge \sous{L}(n),\text{ or }
\ \exists k\ge n,\ \delta_O(k)\ge h(k) \big)\\
\ge &P\big(\delta_I(n)\ge \sous{L}(n)\big)
+P\pare{ \exists k\ge n,\ \delta_O(k)\ge h(k)\ \big|\
\delta_I(n)< \sous{L}(n)}P\big(\delta_I(n)< \sous{L}(n)\big)\\
\ge &1-\Big(1-P\big( \exists k\ge n,\ \delta_O(k)\ge
h(k)\ \big|\ \delta_I(n)< \sous{L}(n)\big)\Big)
P\big(\delta_I(n)< \sous{L}(n)\big).
\end{split}
\end{equation}
We now prove \reff{main-1}. For integer $n$, assume that
$A(b(n)\delta_0)$ is realized.
Let $X_n$ be a Poisson random variable with
parameter $\lambda_n= |{\mathbb B}(0,n+h(n))\backslash{\mathbb B}(0,n)|$.
We realize the cluster $A((b(n)+X_n)\delta_0)$ through three
exploration waves, as explained in Section~\ref{sec-abelian}.
After the first wave with $b(n)$ explorers,
we launch $X_n$ explorers, the green ones, and stop them on
$\Sigma:=\partial {\mathbb B}(0,n-\sous{L}(n))$. Under the event
$\{\delta_I(n)<\sous{L}(n)\}$, note that for $z\in \Sigma$
\[
W_{n-\sous{L}(n)}(A(b(n)\delta_0);X_n\delta_0,z)=
M_{n-\sous{L}(n)}(X_n\delta_0,z).
\]
The configuration of stopped random walks on $\Sigma$
is denoted $\zeta$. Note that $\zeta$ is independent of
$A(b(n)\delta_0)$. For $z\in \Sigma$, we call cov$(z)$ the event that
the $\zeta(z)$ green explorers starting on $z$
produce a cluster $A(\emptyset,\zeta(z)\delta_z)$
which satisfies
\be{one-finger}
A(\emptyset,\zeta(z)\delta_z)\cap {\mathbb B}^c(0,n+4h(n))\not= \emptyset.
\end{equation}
Note that the green explorers contributing to cov$(z)$ start
on the positions of the random walks stopped on $z\in\Sigma$,
which are associated with the green explorers.
Assume, for a moment that when cov$(z)$ happens, there is a tentacle of
$A((b(n)+X_n)\delta_0)$ which protrudes ${\mathbb B}(0,n+4h(n))$.
Assume also that under condition on $X_n$, we have
\be{cond-Xn}
n+4h(n)\ge R_n+h(R_n),\quad\text{where}\quad
R_n=\rho(b(n)+X_n) .
\end{equation}
We would deduce that $\delta_O(R_n)\ge h(R_n)$.
We now proceed through four steps. First, we show that \reff{one-finger}
implies that the final cluster is not inside ${\mathbb B}(0,n+4h(n))$.
Secondly, we estimate the cost of producing a tentacle realizing
cov$(z)$. Then, we establish conditions ensuring \reff{cond-Xn}.
Finally, we show that for an appropriate choice of $\alpha$,
one event cov$(z)$ realizes for some $z\in \Sigma$.
\paragraph{Step 1: Coupling.}
By coupling, it is easy to see that for any subset $\Lambda$,
and $z\in \Sigma$
\be{step-3}
A(\emptyset;\zeta(z)\delta_z)\subset
\Lambda\cup A(\Lambda;\zeta(z)\delta_z)\subset
\Lambda\cup A(\Lambda;\zeta).
\end{equation}
If we denote by $A_2$ the cluster after the second
exploration wave (see Section~\ref{sec-abelian}), then
we have with an equality in law
\be{step-4}
A_2\cup A(A_2; \zeta)=
A\big((b(n)+X_n)\delta_0\big).
\end{equation}
Now, using \reff{step-3}, \reff{step-4},
and \reff{one-finger}, we conclude that
the final cluster is not in ${\mathbb B}(0,n+4h(n))$.
\paragraph{Step 2: Long tentacles.}
To produce cov$(z)$, we first bring a number
of green explorers at $z$ proportional to $h(n)$, and force them to make
a tentacle normal to $\Sigma$ at $z$, with a height $4h(n)+
\sous{L}(n)$. More precisely, draw unit cubes centered on the
points of the sequence
\[
x_n=\big(\|z\|+n\big) \frac{z}{\|z\|}\in {\mathbb R}^d\quad(\text{and}\quad
\|x_n\|=\|z\|+n).
\]
Each such cube contains at least a site of ${\mathbb Z}^d$, say $z_n$.
Note that $\|z_n-z_{n-1}\|\le 2 \sqrt d+1 $, so that we can
exhibit a sequence $\{z=y_1,y_2,\dots,y_N\}$ of nearest neighbors
in ${\mathbb Z}^d$ such that $\|y_N-z\|\ge 4h(n)+\sous{L}(n)$, with
$N\le c(4h(n)+\sous{L}(n))$ for some constant $c$ independent of $n$.
Now, if $M_{n-\sous{L}(n)}(X_n\delta_0,z)\ge N$, and if we launch
the green explorers stopped on $z$ and force them to walk along
the sequence $\{y_1,y_2,\dots,y_N\}$, with the $k$-th explorer
settling on $y_k$, then we realize cov$(z)$, and its probability
is larger than
\[
\pare{\frac{1}{2d}}^{\sum_{k=1}^N k}\ge
\exp\big(-c (4h(n)+\sous{L}(n))^2\big).
\]
\br{rem-d2} In dimension 2, there is a better strategy
to build a tentacle. Since we believe that it yields an estimate
which is not optimal, we do not give the full proof,
but give enough details of the construction so that the interested
reader can easily fill the details.
We first bring a larger number of green
explorers at $z$, about $ \frac{h(n)}{\log^2(h(n))}\times h(n)$.
The probability of so doing is larger than
\be{forcing-many}
\exp\Big(-C \frac{h(n)}{\log^2(h(n))}\log\big(\frac{h(n)}{\log^2(h(n))}
\big) \times h(n)\Big)\ge \exp\Big(-c\frac{h^2(n)}{\log(h(n))}\Big),
\end{equation}
for two positive constants $C,c$. Then, the explorers are
forced to fill sequentially cylindrical compartments
of a telescope-like domain that we now describe. Let $R$ be the integer
part of $4h(n)+\sous{L}(n)$, and
divide $B(z,R)$ into $R$ shells of length $h_1,\dots,h_R$ with
for $i=1,\dots,R$
\be{def-hi}
h_i=\frac{R}{i\times \log(R)}.
\end{equation}
Choose the sequence of $R$ points of ${\mathbb R}^d$
\[
\forall i\in \{1,\dots,R\}\quad
x_i=\big(\|z\|+h_i\big) \frac{z}{\|z\|}
\]
There is $z_i\in \partial {\mathbb B}(z,h_i)$ with $\|z_i-x_i\|\le 2$.
Now, assume we have brought $n_i:=A\pi h_{i+1}^2$ (for a fixed
constant $A$) explorers in ${\mathbb B}(z_i,h_i/4)$.
Then, with a probability larger than $1-1/h_i^2$,
they cover the ball ${\mathbb B}(z_i,h_{i+1})$ by Lemma 1.3 of
\cite{AG2}. Now, if we bring additional explorers in ${\mathbb B}(z_i,h_i/4)$,
they can reach ${\mathbb B}(z_{i+1},h_{i+1}/4)$ with a positive
probability, say $\exp(-\kappa)$:
indeed, they only need to escape a cube-like domain
centered on $z_i$, of side-length $h_{i+1}/4$ on the side which
contains $z_{i+1}$. The cost of the scenario, for which we
only described the $i$-th step, is therefore of order
\be{scenario-d2}
\prod_{i=1}^{R-1}\pare{1-\frac{1}{h_i^2}}\times e^{-\kappa
\sum_{i=2}^R n_i}\times e^{-\kappa
\sum_{i=3}^R n_i}\dots\times e^{-\kappa
n_R}\ge C\exp\big(-\kappa \sum_{i=2}^R (i-1)n_i\big),
\end{equation}
for positive constants $\kappa,C$. Note that with the choice
of $h_i$ in \reff{def-hi}, and $n_i=A\pi h_{i+1}^2$, we have
with $\sous{L}(n)=h(n)$ and for a positive constant $\kappa'$
\[
\sum_{i=2}^R (i-1)n_i\ge \kappa' \frac{h^2(n)}{\log(h(n))}.
\]
Thus, the probability of bringing $h^2(n)/\log^2(n)$ explorers
in $z$, and the probability of building a tentacles of height $5h(n)$
is larger than
\be{order-d2}
\exp\Big(-\kappa_2\frac{h^2(n)}{\log(h(n))}\Big),
\end{equation}
for some positive constant $\kappa_2$.
\end{remark}
\paragraph{Step 3: Bounding $X_n$.}
We impose that $X_n\le 2\lambda_n$.
\be{ie35}
\begin{split}
X_n &\le 2\big|{\mathbb B}(0,n+h(n))\backslash {\mathbb B}(0,n)\big|\\
&\le \big|{\mathbb B}(0,n+2h(n))\backslash {\mathbb B}(0,n)\big|
\Longrightarrow X_n+b(n)\le \big|{\mathbb B}(0,n+2h(n))\big|
\Longrightarrow R_n\le n+2h(n).
\end{split}
\end{equation}
The conclusion of \reff{ie35} implies also that $h(R_n)\le 2h(n)$,
and this implies \reff{cond-Xn}.
Note that since $X_n$ is a Poisson variable of mean $\lambda_n$,
there is a constant $c$, such that
\be{excess-1}
P\big(X_n> 2\lambda_n\big)\le \exp\big(-ch(n)n^{d-1}\big).
\end{equation}
\paragraph{Step 4: Many possible tentacles.}
We summarize Step 1 to Step 3, as establishing that
\be{ie26}
\begin{split}
\acc{ X_n\le 2 \lambda_n,\ \exists z\in \Sigma\quad \text{cov}(z),\
\delta_I(n)<\sous{L}(n)}\subset&
\acc{ \delta_O(R_n)\ge h(R_n),\ \delta_I(n)<\sous{L}(n)}\\
\subset&\acc{\exists k>n,
\delta_O(k)\ge h(k),\ \delta_I(n)<\sous{L}(n)}.
\end{split}
\end{equation}
The key observation now is that $\{ \text{cov}(z),\ z\in \Sigma\}$
are independent Poisson variables which are also independent
of $\delta_I(n)$. Indeed, $\{ \text{cov}(z),\ z\in \Sigma\}$
deals with the walks associated with the green explorers,
whereas $\delta_I(n)$ depends on the $b(n)$ explorers which
we launch first. Taking probability
on both sides of \reff{ie26}, and dividing by $P(\delta_I(n)<\sous{L}(n))$,
we obtain
\be{ie27}
P\big(\exists k>n,
\delta_O(k)\ge h(k)\big| \delta_I(n)<\sous{L}(n)\big)\ge
P\big( X_n\le 2 \lambda_n,\ \exists z\in \Sigma\quad \text{cov}(z)\big).
\end{equation}
Thus,
\be{ie28}
P\big(\exists k>n,
\delta_O(k)\ge h(k)\big| \delta_I(n)<\sous{L}(n)\big)\ge
P\big(\ \exists z\in \Sigma \text{cov}(z)\big)-
P\big( X_n> 2 \lambda_n\big).
\end{equation}
Now, note that for some positive constants $\kappa,\kappa'$
\be{main-3}
\begin{split}
P\pare{\bigcup_{z\in \Sigma} \text{cov}(z)}=& 1-
\prod_{z\in \partial {\mathbb B}}
\Big(1-P\big(\text{cov}(z)
\cap\ \{\zeta(z)>c\big(4h(n)+\sous{L}(n)\big)\}\big)\Big)\\
\ge& 1-\exp\pare{-\kappa' n^{d-1}\exp\pare{-\kappa(h^2(n)+\sous{L}^2(n))}}.
\end{split}
\end{equation}
Now, there is $\alpha>0$ such that for $h(n)\le \alpha\sqrt{\log(n)}$,
and $\sous{L}(n)=h(n)$
\be{step-7}
\lim_{n\to\infty}n^{d-1}\exp\pare{-\kappa(h^2(n)+h(n))}=\infty.
\end{equation}
We now combine \reff{ie28}, \reff{ie26}, \reff{main-3} and
\reff{step-7} to conclude the proof.
In dimension 2, the estimate \reff{main-3} has to be replaced
with \reff{order-d2}.
\section{The inner error}\label{sec-inner}
In this section, we prove Theorem~\ref{theo-lower}.
We show that in the process of going
from a cluster of volume $b(n)$ to one of volume $b(2n)$,
chances tend to one as $n$ tends to infinity,
that there appears a cluster $A$
whose inner error is larger than $\alpha \sqrt{\log(\rho(A))}$
for some positive $\alpha$ independent of $n$.
To do so, we launch many exploration waves, each one is
made up of a Poisson number of explorers.
We proceed inductively.
For positive reals $\alpha,\beta$, to be chosen later,
we set $h(n)=\alpha\sqrt{\log(n)}$, and $\bar{L}(n)
=\beta\sqrt{\log(n)}$. $\bar{L}(n)$ will
refer to an outer radius.
Let $\{{\cal{G}}_n,\ n\ge 0\}$ denotes the natural filtration associated with
the evolution by waves.
First, we launch $b(n)$ explorers.
Assume that explorers of wave $k-1$ have been launched, and are settled.
Knowing ${\cal{G}}_{k-1}$, the size of the $k$-th
wave, denoted $X_k$, is a Poisson variable of parameter
\be{ie1}
\lambda(k)=\big|{\mathbb B}(0,R_{k-1}+2h(R_{k-1}))\backslash{\mathbb B}(0,R_{k-1}))\big|,
\quad\text{where}\quad R_{k-1}=\rho(b(n)+X_1+\dots+X_{k-1}).
\end{equation}
Since $R_{k-1}$ is of order $n$,
each wave fills approximately a peel of width $2h(n)$, and
$n/2h(n)$ waves fill approximately ${\mathbb B}(0,2n)$.
We prove in this section that for an appropriate $\alpha$
\be{ie3}
\lim_{n\to\infty} P\pare{\bigcup_{1\le k<n/2h(n)}
\acc{\delta_I(R_k)> \alpha \sqrt{\log(R_k)}}}=1.
\end{equation}
We now proceed in estimating the probability of observing
a {\it deep} hole after each exploration waves.
We set ${\cal{A}}_k=\acc{\delta_I(R_k)> \alpha \sqrt{\log(R_k)}}$.
\paragraph{On the holes left after wave $k-1$.}
Observe that by definition, on ${\cal{A}}^c_{k-1}$
\[
{\mathbb B}(0,R_{k-1}-h(R_{k-1}))\subset A\big(b(n)+X_1+\dots+X_{k-1}\big),
\]
which implies that
\be{ie4}
\Big({\mathbb B}(0,R_{k-1})\backslash {\mathbb B}(0,R_{k-1}-h(R_{k-1}))\cup
\partial {\mathbb B}(0,R_{k-1})\Big)\cap A\big(b(n)+X_1+\dots+X_{k-1}\big)\not=
\emptyset.
\end{equation}
Choose any $Z_k$ in the intersection of the non-empty set of \reff{ie4},
and note that
\be{ie24}
R_{k-1}-h(R_{k-1})\le \|Z_k\|\le R_{k-1}+1.
\end{equation}
Recall that we have defined $\Sigma(Z_k):=\partial {\mathbb B}(0,\|Z_k\|)$.
We launch the $X_k$ explorers, that we name the {\it green explorers},
and we stop them as they reach $\Sigma(Z_k)$.
The green explorers which settle before reaching $\Sigma(Z_k)$
play no role here, and we bound the number of
green explorers stopped on some region $\Lambda\in \Sigma(Z_k)$,
by the number of corresponding random walks exiting
$\Sigma(Z_k)$ on $\Lambda$. Thus, if we choose
\be{ie5}
\Lambda_k={\mathbb B}(Z_k,\bar L(R_k))\cap \Sigma(Z_k),
\quad\text{and }\quad \Lambda'_k=\Sigma(Z_k)\backslash \Lambda_k,
\end{equation}
and if we denote
\be{ie6}
I_k=\acc{N_{Z_k}(X_k\delta_0,\Lambda_k,\infty)=0,\
N_{Z_k}(X_k\delta_0,\Lambda'_k,7\bar L(R_{k-1}))=0}
\end{equation}
then, on the event $I_k$, green explorers either
exit a ball of radius $R_{k-1}-h(R_{k-1})+7\bar L(R_{k-1})$, or
do not visit $Z_k$. In other words,
\[
I_k\subset \acc{R_k-\delta_I(R_k)< \|Z_k\|}\cup
\acc{R_k+\delta_O(R_k)\ge \|Z_k\|+7\bar L(R_{k-1})}.
\]
In order to conclude that $\{\delta_I(R_k)\ge h(R_k)\}$ or
$\{\delta_O(R_k)\ge \bar L(R_{k})\}$, we need to find
conditions on $X_k$ that guarantee that
\be{cond-Xk}
R_k-\|Z_k\|\ge h(R_k),\quad\text{and}\quad
\|Z_k\|-R_k+7\bar L(R_{k-1})\ge \bar L(R_{k}).
\end{equation}
\paragraph{ Conditions on $X_k$ fulfilling \reff{cond-Xk}.}
We call
\[
{\cal{C}}_k=\acc{2\lambda(k)\ge X_k}\cap \acc{X_k\ge \frac{2}{3}\lambda(k)}.
\]
On the one hand, if $X_k\ge \frac{2}{3}\lambda(k)$, then
\be{ie21}
\begin{split}
X_k &\ge \frac{2}{3}\big|{\mathbb B}(0,R_{k-1}+2h(R_{k-1}))\backslash {\mathbb B}(0,R_{k-1})\big|\\
&\ge \big|{\mathbb B}(0,R_{k-1}+\frac{4}{3}h(R_{k-1}))\backslash {\mathbb B}(0,R_{k-1})\big|
\Longrightarrow R_k\ge R_{k-1}+\frac{4}{3} h(R_{k-1}).
\end{split}
\end{equation}
On the other hand, if $X_k\le 2 \lambda_k$, then
\be{ie22}
\begin{split}
X_k &\le 2\big|{\mathbb B}(0,R_{k-1}+2h(R_{k-1}))\backslash {\mathbb B}(0,R_{k-1})\big|\\
&\le \big|{\mathbb B}(0,R_{k-1}+4h(R_{k-1}))\backslash {\mathbb B}(0,R_{k-1})\big|
\Longrightarrow R_k\le R_{k-1}+4h(R_{k-1}).
\end{split}
\end{equation}
Now, for $x$ large enough, the following implication is obvious
\be{obvious-1}
x\le y+4h(y)\Longrightarrow h(x)\le \frac{4}{3}h(y)-1
,\quad\text{and}\quad \bar L(x)\le 2 \bar L(y).
\end{equation}
If $n$ is large enough, \reff{obvious-1} and \reff{ie22} imply that
$h(R_k)\le \frac{4}{3}h(R_{k-1})-1$, which in turn, with \reff{ie21},
yields
\be{ie23}
R_k\ge R_{k-1}-1+h(R_k).
\end{equation}
Also, $\bar L(R_k)\le 2 \bar L(R_{k-1})$, $\bar L(R_{k-1})\ge
h(R_{k-1})$, in combination with \reff{ie22} imply that
\be{ie25}
R_{k-1}-h(R_{k-1})+7\bar L(R_{k-1})-\bar L(R_{k})\ge
R_{k-1}+4h(R_{k-1})\ge R_k.
\end{equation}
Thus, if ${\cal{C}}_k\cap {\cal{A}}^c_{k-1}$ holds, then
\reff{ie24}, \reff{ie23} and \reff{ie25} imply that
conditions \reff{cond-Xk} holds.
\paragraph{On a deep hole in one shell.}
We choose an integer $k<n/2h(n)$.
We have seen that knowing ${\cal{A}}^c_{k-1}$
\be{ie10}
I_k\cap {\cal{C}}_k\subset {\cal{A}}_k \cup \acc{\delta_0(R_k)\ge \bar L(R_k)}.
\end{equation}
Taking conditional probabilities on both sides of \reff{ie10},
we obtain,
\be{ie20}
\begin{split}
\hbox{ 1\hskip -3pt I}_{{\cal{A}}^c_{k-1}}
P\big({\cal{A}}^c_k\cap{\cal{C}}_k\ \big|{\cal{G}}_{k-1}\big)=&
\hbox{ 1\hskip -3pt I}_{{\cal{A}}^c_{k-1}}\big(P({\cal{C}}_k)-P({\cal{A}}_k\cap {\cal{C}}_k\big|{\cal{G}}_{k-1})\big)\\
\le &
\hbox{ 1\hskip -3pt I}_{{\cal{A}}^c_{k-1}}\Big(
P({\cal{C}}_k)-P(I_k\cap {\cal{C}}_k\big|{\cal{G}}_{k-1})+
P\big(\acc{\delta_0(R_k)\ge \bar L(R_k)}|{\cal{G}}_{k-1}\big)\Big)\\
\le &P(\acc{\delta_0(R_k)\ge \bar L(R_k)}|{\cal{G}}_{k-1})
+\hbox{ 1\hskip -3pt I}_{{\cal{A}}^c_{k-1}}\big(1-P(I_k\big|{\cal{G}}_{k-1})\big).
\end{split}
\end{equation}
Now, we invoke Lemma~\ref{lem-main}
with $\|z\|,R$ and $\lambda$ respectively of order
$n,\bar L(n)$, and $h(n) n^{d-1}$. As a consequence,
we have on ${\cal{C}}_k\cap {\cal{A}}^c_{k-1}$ for a constant $\kappa$,
\be{ie-main}
\inf_{k\le n/2h(n)} P(I_k\ |\ {\cal{G}}_{k-1})\ge \exp(-\kappa h(n) \bar L(n)).
\end{equation}
If we denote $N$ for the integer part of $n/2h(n)$, and proceed
inductively, we obtain
\[
\begin{split}
P\big(\cup_{k\le N} {\cal{A}}_k\big)&
-P\big(\cap_{k\le N} {\cal{C}}_k\big)\ge-
P\big(\forall k\le N,\ {\cal{A}}^c_k\cap{\cal{C}}_k\big)\\
&\ge -E\cro{\hbox{ 1\hskip -3pt I}_{\forall k<N,\ {\cal{A}}^c_k\cap{\cal{C}}_k}
P\big({\cal{A}}^c_N\cap{\cal{C}}_N\ \big|{\cal{G}}_{N-1}\big)} \\
&\ge -E\cro{\hbox{ 1\hskip -3pt I}_{\forall k<N,\ {\cal{A}}^c_k\cap{\cal{C}}_k}
\big(P(\acc{\delta_0(R_N)\ge \bar L(R_N)}|{\cal{G}}_{N-1})
+1-P(I_N\big|{\cal{G}}_{N-1})}\big)\\
&\ge -P(\acc{\delta_0(R_N)\ge \bar L(R_N)})-(1-
\exp(-\kappa h(n) \bar L(n)))P\pare{\forall k<N,\ {\cal{A}}^c_k\cap{\cal{C}}_k}\\
&\ge -\sum_{k\le N} P(\acc{\delta_0(R_k)\ge \bar L(R_k)})-
(1-\exp(-\kappa h(n) \bar L(n)))^{N}.
\end{split}
\]
Thus,
\be{ie11}
P\big(\cup_{k\le N} {\cal{A}}_k\big)\ge 1-
\sum_{k\le N} \Big(P\big(\acc{\delta_0(R_k)\ge \bar L(R_k)}\big)+
P({\cal{C}}_k^c)\Big)-(1-\exp(-\kappa h(n) \bar L(n)))^{N}.
\end{equation}
Now, we have established in \cite{AG2}, that for $\beta$ large
enough, the probability
of $\{\delta_0(R_k)\ge \bar L(R_k)\}$ decays faster than any power
in $n$, whereas the fact that $X_k$ is Poisson implies that for some
constant $c$, we have $P({\cal{C}}_k^c)\le \exp(-c h(n)n^{d-1})$. The last term
on the last display of \reff{ie11} tends to 0 if
\be{cond-h}
\lim_{n\to\infty} \frac{n}{2h(n)}\exp(-\kappa h(n) \bar L(n))=
\infty.
\end{equation}
In dimension 3 or more, \reff{cond-h} holds
for $\alpha$ small enough.
\section{Proof of Proposition~\ref{prop-AG}}\label{sec-last}
The proof is a direct corollary of formula (3.11) of \cite{AG2}.
We consider actually {\it tiles} of size 1, that is site of ${\mathbb Z}^d$.
Inequality (3.8) of \cite{AG2} shows that for some constant $c_d$
(depending only on dimension) and $R=\|z\|<n$, we have
\be{ag2-1}
E[W_R(\emptyset, N\delta_0,z)]\ge c_d (n-\|z\|).
\end{equation}
Inequality (3.10) of \cite{AG2} is written a little differently as
\be{outer-1}
P\pare{ W_{R}(\emptyset, N\delta_0,z)=0}\le \left\{ \begin{array}{ll}
\exp\pare{-\lambda \kappa_2 (n-\|z\|) +
\lambda^2 c'_2 \log(n)}& \mbox{ for } d=2 \, , \\
\exp\pare{-\lambda \kappa_d (n-\|z\|) +
\lambda^2 c'_d }& \mbox{ for } d\ge 3 \, .
\end{array} \right.
\end{equation}
As we optimize \reff{outer-1} in $\lambda>0$, we obtain
\reff{AG-directional}.
|
\section{Introduction}
Although known since the time of Lifshitz's work on the subject
\cite{lifshitz61}, repulsive Casimir forces have recently received
serious scrutiny \cite{Levin:2010zz}. Experimental confirmation
of the repulsion that occurs when dielectric surfaces are separated
by a liquid with an intermediate value of the dielectric constant
has appeared \cite{capasso09}, although this seems devoid of much
practical application. The context of our work is the
considerable interest in utilizing the quantum vacuum force or
the Casimir effect in nanotechnology employing mesoscopic objects
\cite{Rodriguez:2010zz}.
The first repulsive Casimir stress in vacuum was found by Boyer
\cite{boyer68}, who discovered the still surprising fact that
the Casimir self-energy of a perfectly conducting spherical
shell is positive. (This has become somewhat less mysterious,
since the phenomenon is part of a general pattern
\cite{bender,milton,abalo1,abalo2}.) Boyer later observed that
a perfect electrical conductor and a perfect magnetic conductor
repel \cite{boyer74}, but this also seems beyond reach, since the unusual
electrical properties must be exhibited over a wide frequency range. The
analogous effect for metamaterials also seem impracticable \cite{McCauley}.
Thus it was a significant advance when Levin et al.\
showed examples of repulsion between conducting objects, in particular
between an elongated cylinder above a conducting plane with a circular
aperture \cite{Levin:2010zz}.
(See also Ref.~\cite{maghrebi}.)
They computed the quantum vacuum forces between conducting objects, by using
impressive numerical finite-difference time-domain and
boundary-element methods.
We subsequently showed \cite{Milton:2011ni} that repulsive Casimir-Polder
forces between anisotropic atoms and a conducting half-plane, and even
between such an atom and a conducting wedge of rather large opening angle,
could be achieved. Of course, we must be careful to explain what we mean
by repulsion: the total force on the atom is attractive, but the component
of the force perpendicular to the symmetry axis of the conductor changes
sign when the atom is sufficiently close to that axis. This is the
only component that survives in the case of an aperture in a plane, so
our analytic calculation provided a counterpart to the numerical work
of Ref.~\cite{Levin:2010zz}.
In this paper we give some further examples. After demonstrating, in
Sec.~\ref{sec2}, that Casimir-Polder repulsion
between two atoms requires that both be sufficiently
anisotropic, we show in Sec.~\ref{sec3}
that the force between one such atom and a conducting
cylinder is repulsive for motion confined to a perpendicular line not intersecting with
the cylinder, provided the line is sufficiently far from the cylinder.
The analogous effect does not occur for a spherical conductor
(Sec.~\ref{sec4}), as one
might suspect since at large distances such a sphere looks like an isotropic
atom. The classical interaction between a dipole and a conducting ellipsoid
polarized by an external field is examined in Sec.~\ref{sec5}, which, as
expected, yields a repulsive region. In contrast, in Sec.~\ref{sec6}, we examine the
Casimir-Polder interaction
of an anisotropic atom with an anisotropic dielectric half-space, but this fails
to reveal any repulsive regime.
In this paper we set $\hbar=c=1$, and all results are expressed in Gaussian
units except that Heaviside-Lorentz units are used for Green's dyadics.
\section{Casimir-Polder repulsion between atoms}
\label{sec2}
The interaction between two polarizable atoms, described by general
polarizabilities $\bm{\alpha}_{1,2}$, with the relative separation
vector given by $\mathbf{r}$ is \cite{CP1,CP2}
\begin{equation}
U_{\rm CP}=-\frac1{4\pi r^7}\left[\frac{13}2\Tr\bm{\alpha}_1\cdot
\bm{\alpha}_2-28\Tr(\bm{\alpha}_1\cdot\mathbf{\hat r})(\bm{\alpha}_2
\cdot\mathbf{\hat r})+\frac{63}2(\mathbf{\hat r}\cdot\bm{\alpha}_1
\cdot\mathbf{\hat r})(\mathbf{\hat r}\cdot\bm{\alpha}_2
\cdot\mathbf{\hat r})\right].\label{generalcp}
\end{equation}
This formula is easily rederived by the multiple scattering
technique as explained in Ref.~\cite{brevikfest}.
This reduces, in the isotropic case, $\bm{\alpha}_i=\alpha_i\bm{1}$,
to the usual Casimir-Polder (CP) energy,
$U_{\rm CP}=-\frac{23}{4\pi r^7}\alpha_1\alpha_2$.
Suppose the two atoms are only polarizable in perpendicular directions,
$\bm{\alpha}_1=\alpha_1\mathbf{\hat z \hat z}$,
$\bm{\alpha}_2=\alpha_2\mathbf{\hat x \hat x}$. Choose atom 2 to be at the origin.
The configuration is shown in Fig.~\ref{pol-atoms}.
\begin{figure}
\begin{center}
\includegraphics{pol-atoms.eps}
\caption{\label{pol-atoms} Casimir-Polder interaction between two atoms
of polarizability $\bm{\alpha}_1$ and $\bm{\alpha}_2$ separated by a distance $r$.
Atom 1 is predominantly polarizable in the $z$ direction, while atom 2
is predominantly polarizable in the $x$ direction. The force on atom 1
in the $z$ direction becomes repulsive sufficiently close to the polarization
axis of atom 2 provided both atoms are sufficiently anisotropic.}
\end{center}
\end{figure}
Then, in terms of the polar angle $\cos\theta=z/r$, the $z$-component of
the force on atom 1 is
\begin{equation}
F_z=-\frac{63}{8\pi}\frac{\alpha_1\alpha_2}{x^8}\sin^{10}\theta\cos\theta
(9-11\sin^2\theta).
\end{equation}
In this paper, we are considering motion for fixed $x=r\sin\theta$,
in the $y=0$ plane.
Evidently, the force is attractive at large distances, vanishing
as $\theta\to0$, but it must change sign at small values of $z$ for fixed $x$,
since the energy also vanishes as $\theta\to\pi/2$. The force component
in the $z$ direction vanishes when $\sin\theta=3/\sqrt{11}$ or $\theta=1.130$
or 25$^\circ$ from the $x$ axis.\footnote{After the first version of this paper
was prepared, Ref.~\cite{SS} appeared, which rederived these results, and then
went on to extend the calculation to Casimir-Polder repulsion by an anisotropic
dilute dielectric sheet with a circular aperture. The authors quite correctly
point out that the statement in Ref.~\cite{Milton:2011ni} that no repulsion is possible in
the weak-coupling regime is erroneous.}
No repulsion occurs if one of the atoms is isotropically polarizable.
If both have cylindrically symmetric anisotropies, but with respect
to perpendicular axes,
\begin{equation}
\bm{\alpha}_1=(1-\gamma_1)\alpha_1\mathbf{\hat z \hat z}+\gamma_1\alpha_1\bm{1},\quad
\bm{\alpha}_2=(1-\gamma_2)\alpha_2\mathbf{\hat x \hat x}+\gamma_2\alpha_2\bm{1},
\end{equation}
it is easy to check that if both are sufficiently anisotropic repulsion will occur.
For example, if $\gamma_1=\gamma_2$ repulsion in the $z$ direction will take place close to the plane
$z=0$ if $\gamma\le0.26$.
\section{Repulsion of an atom by a conducting cylinder}
\label{sec3}
Now we turn to the Casimir-Polder (CP) interaction between a polarizable
body (``atom'') and a macroscopic body. That interaction is generally given by
\begin{equation}
E_{\rm CP}=-\int_{-\infty}^\infty d\zeta\tr \bm{\alpha}\cdot\bm{\Gamma}(\mathbf{r,r}),
\end{equation}
where $\mathbf{r}$ is the position of the atom and $\zeta$ is
the imaginary frequency,
in terms of the polarizability of the atom $\bm{\alpha}$ and
the Green's dyadic due to the macroscopic body, which for
a body characterized by a permittivity $\varepsilon$
satisfies the differential equation
\begin{equation}
\left(\frac1{\omega^2}\bm{\nabla}\times\bm{\nabla}\times
-\bm{1}\varepsilon(\mathbf{r})\right)\cdot\bm{\Gamma}(\mathbf{
r,r'})=\bm{1}\delta(\mathbf{r-r'}).\label{gdif}
\end{equation} In this paper, except for Sec.~\ref{sec6},
we will consider perfect conducting boundaries $S$
immersed in vacuum, in which case
we need to solve this equation with $\varepsilon=1$ for $\bm{\Gamma}$
subject to the boundary conditions
$\mathbf{\hat n}\times \bm{\Gamma}(\mathbf{r,r'})\bigg|_{\mathbf{r}\in S}=0$,
where $\mathbf{\hat n}$ is the normal to the surface of the conductor,
which just states that the tangential components of the
electric field must vanish on the conductor.
Let us henceforth assume that the polarizability has negligible frequency dependence
(static approximation), and, in order to maximize the repulsive effect, the atom
is only polarizable in the $z$ direction, the direction of the trajectory
(assumed not to intersect the cylinder),
in which case the quantity we need to compute for
a conducting cylinder of radius $a$ is given by \cite{bezerra}
\begin{eqnarray}
\int_{-\infty}^\infty \frac{d\zeta}{2\pi}\Gamma_{zz}(r,\theta)&=
&\sum_{m=-\infty}^\infty
\int_0^\infty\frac{d\kappa}{(2\pi)^3}\frac\pi{2a}
\frac1{K_m(\kappa a)K'_m(\kappa a)}
\bigg\{\frac{m^2}{r^2}K_m^2(\kappa r)+\kappa^2K_m^{\prime2}(\kappa r)\nonumber\\
&&\quad\mbox{}-\cos2\theta \kappa a[I_m(\kappa a)K_m(\kappa a)]'
\left(-\frac{m^2}{r^2}
K_m^2(\kappa r)+\kappa^2K_m^{\prime 2}(\kappa r)\right)\bigg\}.
\end{eqnarray}
The geometry we are considering is illustrated in Fig.~\ref{fig-cyl-atom}.
\begin{figure}
\begin{center}
\includegraphics{cyl-atom.eps}
\caption{\label{fig-cyl-atom} Interaction between an anisotropically polarizable
atom and a conducting cylinder of radius $a$. The force on the atom along a line
which does not intersect the cylinder is considered. If the atom is only
polarizable in that direction, and the line lies sufficiently far from the
cylinder, the force component along the line changes sign near the point of
closest approach.}
\end{center}
\end{figure}
It gives greater insight to give the transverse electric (TE)
and transverse magnetic (TM) contributions to the CP energy:
\begin{subequations}
\begin{eqnarray}
E^{\rm TE}_{\rm CP}&=&-\frac{\alpha_{zz}}{4\pi}\sum_{m=-\infty}^\infty
\int_0^\infty d\kappa\,\kappa
\frac{I'_m(\kappa a)}{K'_m(\kappa a)}\left[\frac{\cos^2\theta}{r^2} m^2K_m^2
(\kappa r)
+\kappa^2\sin^2\theta K_m^{\prime 2}(\kappa r)\right],\\
E^{\rm TM}_{\rm CP}&=&\frac{\alpha_{zz}}{4\pi}\sum_{m=-\infty}^\infty
\int_0^\infty d\kappa\,\kappa
\frac{I_m(\kappa a)}{K_m(\kappa a)}\left[\frac{\sin^2\theta}{r^2} m^2K_m^2
(\kappa r)+\kappa^2\cos^2\theta K_m^{\prime 2}(\kappa r)\right].
\end{eqnarray}
\end{subequations}
The distance of the atom from the center of the
cylinder is $r=R/\sin\theta$, where $R$ is the distance of
closest approach and $\theta$ is the polar angle, which ranges from
0 when the atom is at infinity to $\pi/2$ when the atom is closest to
the cylinder.
At large distances, the CP force is dominated by the $m=0$ term in the
energy sum. Figure \ref{fig1} shows that for $m=0$ the TM mode dominates
except near the position of closest approach, where only the TE mode
is nonzero. This indicates that there is a region of repulsion near
$\theta=\pi/2$, since the total energy has a minimum for small
$\psi=\pi/2-\theta$. This effect is partially washed out by including higher $m$ modes,
as seen in Fig.~\ref{fig2}, which shows the effect of including the first 5
$m$ values. But the repulsion goes away if the line of motion passes
too close to the cylinder. Numerically, we have found that to have
repulsion close to the plane of closest approach requires that
$a/R<0.15$.
\begin{figure}
\begin{center}
\includegraphics{new-cyl-cp-rep.eps}
\caption{\label{fig1} $m=0$ contributions to the Casimir-Polder energy
between an anisotropic atom and a conducting cylinder. The (generally) lowest curve
(blue) is the TE contribution, the second (magenta) is the TM contribution, and the top
curve (yellow) is the total CP energy. In this case, the distance of closest
approach of the atom is taken to be 10 times the radius of the cylinder.
The energy $E$ is plotted as a function of $\psi=\pi/2-\theta$.}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics{new-cyl-cp-rep2.eps}
\caption{\label{fig2} The CP energy between an anisotropic atom
and a conducting cylinder. Plotted is the total CP energy, the upper curve
for the distance of closest approach $R$ being 5 times the cylinder
radius $a$, the lower curve for the distance of closest approach 10 times the
radius. The curves move up slightly as more $m$ terms are included,
but have completely converged by the time $m=3$ is included.
Repulsion is clearly observed when $R/a=10$, but not for $R/a=5$.}
\end{center}
\end{figure}
\section{CP interaction between atom and conducting sphere}
\label{sec4}
It is straightforward to derive the TE and TM contributions
for the interaction between a completely anisotropic
atom and a conducting sphere as
\begin{subequations}
\begin{eqnarray}
E^{\rm TM}&=&\frac{\alpha_{zz}}{2\pi R^4}\cos^4\theta\sum_{l=1}^\infty
(2l+1)\int_0^\infty dx \,g_l(x),\\
E^{\rm TE}&=&\frac{\alpha_{zz}}{4\pi R^4}\cos^6\theta\sum_{l=1}^\infty
(2l+1)\int_0^\infty dx \,f_l(x),
\end{eqnarray}
\end{subequations}
where
\begin{subequations}
\begin{eqnarray}
g_l(x)&=&x\frac{s_l'(xa\cos\theta/R)}{e_l'(xa\cos\theta/R)}\left[\frac12\cos^2
\theta e_l^{\prime2}(x)+\frac{l(l+1)\sin^2\theta e_l^2(x)}{x^2}\right],\\
f_l(x)&=&x\frac{s_l(xa\cos\theta/R)}{e_l(xa\cos\theta/R)}e_l^2(x),
\end{eqnarray}
\end{subequations}
where the modified Riccati-Bessel functions are
\begin{equation}
s_l(x)=\sqrt{\frac{\pi x}2}I_{l+1/2}(x),\quad
e_l(x)=\sqrt{\frac{2 x}\pi}K_{l+1/2}(x).\quad
\end{equation}
We expect in the case of a sphere not to see Casimir repulsion at large
distances. The reason is that far from the sphere it appears to be an
isotropic atom, which, as we have seen above will not give a
repulsive force on another completely anisotropic atom. Indeed, far
from the sphere we can replace the Bessel functions of argument $xa/r$
by their leading small argument approximations and we easily find
\begin{subequations}
\begin{equation}
E^{\rm TM}\sim \frac{\alpha_{zz} a^3}{4\pi r^7}(13+7\sin^2\theta),
\quad a/r\to 0.\label{TMcp}
\end{equation}
The TE mode contributes
\begin{equation}
E^{\rm TE}\sim \frac{\alpha_{zz}a^3}{4\pi r^7}\frac74\cos^2\theta,
\quad a/r\to0.\label{cpte}
\end{equation}
\end{subequations}
We see here the expected isotropic electric
polarizability of a conducting sphere
$\alpha_{{\rm sp},E}=\bm{1} a^3$.
We note that the TM result (\ref{TMcp}) coincides with the result obtained
from Eq.~(\ref{generalcp}). The TE contribution is, in fact, the coupling
between the electric polarizability of the atom and the magnetic polarizability
of the sphere
$\alpha_{{\rm sp},M}=-\frac{a^3}2\bm{1}$ \cite{embook}.
To see this, we first remind the reader of the CP interaction between
isotropic atoms possessing both electric and magnetic polarizabilities
\cite{feinberg},
\begin{equation}
U_{\rm CP}=-\frac{23}{4\pi r^7}(\alpha_1^E\alpha_2^E+\alpha_1^M\alpha_2^M)
+\frac7{4\pi r^7}(\alpha_1^E\alpha_2^M+\alpha_1^M\alpha_2^E).\label{totalcp}
\end{equation}
When the atoms are not isotropic it is easy to deduce the generalization
of this, using the methods described in
Ref.~\cite{brevikfest}, starting from the
multiple-scattering coupling term between electric
and magnetic dyadics,
\begin{equation}
E_{\rm em}=-\frac{i}2\Tr\ln\left(1+\bm{\Phi}_0 \mathbf{T}_1^E\cdot\bm{\Phi}_0
\mathbf{T}_2^M\right)\approx
-\frac{i}2\Tr\bm{\Phi}_0\cdot\mathbf{V}_1^E\bm{\Phi}_0^e\cdot \mathbf{V}_2^M,
\end{equation}
where the last form reflects weak coupling, and we are considering the
interaction between one object having purely electric susceptibility and
a second object having purely magnetic susceptibility, so
\begin{equation}
\mathbf{V}_1^E=4\pi \bm{\alpha}^E_1\delta(\mathbf{r-r_1}),\quad
\mathbf{V}_2^M=4\pi \bm{\alpha}^M_2\delta(\mathbf{r-r_2}).
\end{equation}
This formula is expressed in terms of the magnetic Green's dyadic,
\begin{equation}
\bm{\Phi}_0=-\frac{\zeta^2}{4\pi R^3}\mathbf{R}\times(|\zeta| R+1)e^{-|\zeta|R}.
\end{equation}
Then, an immediate calculation yields the electric-magnetic CP interaction
\begin{equation}
U_{\rm CP,EM}=\frac7{8\pi R^7}\tr (\mathbf{\hat R}\times\bm{\alpha}^E)
(\mathbf{\hat R}\times\bm{\alpha}^M),
\end{equation}
which indeed for isotropic polarizabilities
gives the second term in Eq.~(\ref{totalcp}). The result
(\ref{cpte}) is now an immediate consequence for a conducting sphere
interacting with an atom only polarizable in the $z$ direction.
Evidently, no repulsion can occur in this CP limit where the conducting sphere
is regarded as an
isotropically polarizable atom. In fact, numerical evaluation
shows no repulsion occurs at any separation distance between the
sphere and the atom.
\section{ Electrostatic force between a conducting ellipsoid and a dipole}
\label{sec5}
In this section we return, for heuristic reasons, to the electrostatic situation of the
interaction between a fixed dipole and a conducting body. Such have been given
considerable attention lately \cite{Levin:2010zz,Milton:2011ni,lj}. Here we
consider the interaction between a perfectly conducting ellipsoid polarized by a
constant electric field and a fixed dipole. The polarization of the ellipsoid
by the dipole is neglected at this stage. This is a much simpler calculation than
the more interesting one of the interaction between a dipole and a ellipsoid, but we
justify the inclusion of the details of the simpler calculation here because it allows
us to approach the
complexity of the full calculation. Elsewhere, we will present that calculation and
the corresponding quantum Casimir-Polder calculation, building on the work of
Ref.~\cite{Graham:2011ta}.
\subsection{Ellipsoidal coordinates}
Consider first a conducting uncharged solid ellipsoid with semiaxes $a>b>c$, centered at the origin $x=y=z=0$. The semiaxis $c$ lies along the $z$ axis. The electrostatic potential $\phi$ in the external region can be described in terms of ellipsoidal coordinates
$\xi, \eta, \zeta$, corresponding to solutions for $u$ of the cubic equation
\begin{equation}
\frac{x^2}{a^2+u}+\frac{y^2}{b^2+u}+\frac{z^2}{c^2+u}=1. \label{1}
\end{equation}
The coordinate intervals are in general
\begin{equation}
\infty >\xi \geq -c^2, \quad -c^2 \geq \eta \geq -b^2, \quad -b^2\geq \zeta \geq -a^2. \label{2}
\end{equation}
We will henceforth assume axial symmetry around the $z$ axis. In that case, $b \rightarrow a, \,\zeta \rightarrow -a^2 $, and the ellipsoidal coordinates $\xi, \eta, \zeta$ reduce to oblate spheroidal coordinates $\xi$ and $\eta$ restricted to the intervals
\begin{equation}
\infty > \xi \geq -c^2, \quad -c^2 \geq \eta \geq -a^2. \label{3}
\end{equation}
If $\rho=\sqrt{x^2+y^2}$ denotes the horizontal radius in the plane $z=$ constant, the cubic equation (\ref{1}) reduces to the quadratic equation
\begin{equation}
u^2-(\rho^2-a^2-c^2+z^2)u-(\rho^2-a^2)c^2-z^2a^2=0 \label{4}
\end{equation}
for $u=(\xi, \eta)$. The solution for $u=\xi$ corresponds to the positive square root:
\begin{equation}
\xi=\frac{1}{2}(\rho^2-a^2-c^2+z^2)+\frac{1}{2}\sqrt{(\rho^2-a^2+c^2)^2+z^2(2\rho^2+2a^2-2c^2+z^2)}. \label{5}
\end{equation}
At the surface of the ellipsoid, $\xi=0$, whereas in the external region, $\xi>0$. Note that in the $xy$ plane ($z=0$) the expression for $\xi$ simplifies to $\xi=\rho^2-a^2$, when
$\rho>a$. The solution for $u=\eta$ corresponds to the same expression (\ref{5}) but with the negative square root.
Surfaces of constant $\xi$ and $\eta$ are oblate spheroids and hyperboloids of revolution, the surfaces intersecting orthogonally. On the symmetry axis $\rho=0$ one has $\xi=-c^2+z^2, \, \eta=-a^2$. The relations between $\xi, \eta$ and $z, \rho$ are
\begin{equation}
z=\pm \sqrt{ \frac{(\xi+c^2)(\eta+c^2)}{c^2-a^2}}, \quad \rho=\sqrt{\frac{(\xi+a^2)(\eta+a^2)}{a^2-c^2}}. \label{6}
\end{equation}
We will henceforth be concerned with the half-space $z\geq 0$ only.
\subsection{Ellipsoid situated in a uniform electric field}
Assume now that the ellipsoid is placed in a uniform electric field ${\bf E}_0$, directed along the $z$ axis. We take the electrostatic potential $\phi$ to be zero on the ellipsoid surface. With quantities $R_\xi$ and $R_\eta$ defined as
\begin{equation}
R_\xi=(\xi+a^2)\sqrt{\xi+c^2}, \quad R_\eta=(\eta+a^2)\sqrt{\eta+c^2}, \label{7}
\end{equation}
the Laplace equation in the external region $\xi \geq 0$ can be written as
\begin{equation}
\nabla^2 \phi \equiv \frac{4}{\xi-\eta}\left[ \frac{R_\xi}{\xi+a^2}\frac{\partial}{\partial\xi}\left(R_\xi\frac{\partial
\phi}{\partial \xi}\right)- \frac{R_\eta}{\eta+a^2} \frac{\partial}{\partial \eta} \left(R_\eta \frac{\partial
\phi}{\partial \eta}\right)\right]=0. \label{8}
\end{equation}
The potential due solely to ${\bf E}_0$ is
\begin{equation}
\phi_0=-E_0 z, \label{9}
\end{equation}
and we write the full potential $\phi$ in the form
\begin{equation}
\phi=\phi_0[1+F(\xi)], \label{10}
\end{equation}
so that $\phi_0F$ denotes the modification due to the ellipsoid. The boundary condition at the surface is $F(0)=-1$.
Inserting Eq.~(\ref{10}) into Eq.~(\ref{8}) we find the following equation for $F$,
\begin{equation}
\frac{d^2 F}{d\xi^2}+\frac{dF}{d\xi}\frac{d}{d\xi}\ln \left[ R_\xi (\xi+c^2)\right]=0. \label{11}
\end{equation}
The solution can be written as
\begin{equation}
\phi=\phi_0\left[ 1-\frac{\int_\xi^\infty \frac{ds}{(s+c^2)R_s}}{\int_0^\infty \frac{ds}{(s+c^2)R_s}} \right].
\label{12}
\end{equation}
We can also express the solution in terms of the incomplete beta function,
defined as
\begin{equation}
B_x(\alpha, \beta)=\int_0^x t^{\alpha-1}(1-t)^{\beta-1} dt. \label{13}
\end{equation}
Some manipulation yields
\begin{equation}
\int_\xi^\infty \frac{ds}{(s+c^2)R_s}=\frac{1}{(a^2-c^2)^{3/2}}
B_{(a^2-c^2)/(\xi+a^2)}\left(\frac{3}{2}, -\frac{1}{2}\right), \label{14}
\end{equation}
and so we can write the final answer for the potential as
\begin{equation}
\phi=\phi_0\left[ 1-\frac{B_{(a^2-c^2)/(\xi+a^2)}\left(
\frac{3}{2}, -\frac{1}{2}\right)}{B_{1-c^2/a^2}\left(
\frac{3}{2}, -\frac{1}{2}\right)} \right]. \label{15}
\end{equation}
For small values of $x$ the following expansion may be useful,
\begin{equation}
B_x(\alpha, \beta)=\frac{x^\alpha}{\alpha}(1-x)^\beta \left[1+
\sum_{n=0}^\infty \frac{B(\alpha+1,n+1)}{B(\alpha+\beta,n+1)}x^{n+1}\right],
\label{16}
\end{equation}
where $B(\alpha,\beta)=\Gamma(\alpha)\Gamma(\beta)/\Gamma(\alpha+\beta)$
is the complete beta function. In our case, the limit $x \ll 1$
corresponds to the minor semiaxis $c$ being only slightly less than the
major semiaxis $a$.
In the following, we shall need the expression for the $z$ component of the
electric field, $E_z=-\partial \phi/\partial z$, at an arbitrary point
$(\rho,z)$ in the exterior region. It is here convenient first to
differentiate the relation (\ref{4}) ($u=\xi$) with respect to $z$,
keeping $\rho$ constant, to obtain
\begin{equation}
\left(\frac{\partial \xi}{\partial z}\right)_\rho=\frac{2(\xi+a^2)}
{\xi-\eta}\sqrt{\frac{(\xi+c^2)
(\eta+c^2)}{c^2-a^2}}. \label{17}
\end{equation}
With $x=(a^2-c^2)/(\xi+a^2)$ we have
\begin{equation}
\frac{\partial B_x \left(\frac{3}{2},-\frac{1}{2}\right)}
{\partial z}=\frac{\partial \xi}{\partial z}\,\frac{\partial x}
{\partial \xi}\,\frac{\partial B_x\left(\frac{3}{2},-\frac{1}{2}\right)}
{\partial x}=2\frac{(a^2-c^2)}{(\xi+c^2)(\xi-\eta)}(-\eta-c^2)^{1/2}.
\label{18}
\end{equation}
Then, from Eq.~(\ref{15}),
\begin{equation}
E_z=E_0\left[ 1-\frac{B_{(a^2-c^2)/(\xi+a^2)}\left(\frac{3}{2},-\frac{1}{2}
\right)}{B_{1-c^2/a^2}\left(\frac{3}{2}, -\frac{1}{2}\right)}
-\frac{2(a^2-c^2)^{1/2}(\xi+c^2)^{-1/2}(\eta+c^2)}{B_{1-c^2/a^2}
\left(\frac{3}{2}, -\frac{1}{2}\right)}\frac1{\xi-\eta} \right]. \label{19}
\end{equation}
For large values of $z$ and arbitrary $\rho$ the influence from the
ellipsoid must evidently fade away, $E_z \rightarrow E_0$.
In the $xy$ plane where $z=0, \xi+a^2=\rho^2$, $\eta+c^2=0$, we have
\begin{equation}
E_z(z=0)=E_0\left[ 1-\frac{B_{(a^2-c^2)/\rho^2}\left(\frac{3}{2},
-\frac{1}{2}\right)}{B_{1-c^2/a^2}\left(\frac{3}{2}, -\frac{1}{2}\right)}
\right]. \label{20}
\end{equation}
When $\rho=a$ (on the surface), $E_z(z=0)=0$ as it should.
\subsection{Force on a dipole}
Assume now that a dipole ${\bf p}=p_z \mathbf{\hat z}$
is situated at rest in the position $(\rho, z)$.
The dipole is taken to be polarized in the $z$
direction only. The value of $z$ $( \geq 0)$ is arbitrary,
whereas the value of $\rho$ is assumed constant. Thus,
writing $\rho=a+L$, $L$ is the constant horizontal distance between
the dipole and the edge of the ellipsoid. The force $F_z$ on the dipole is
\begin{equation}
F_z=\nabla_z({\bf p\cdot E}) =p_z\frac{\partial E_z}{\partial z}. \label{21}
\end{equation}
Note that we are ignoring the polarization of the ellipsoid by the field
of the dipole; the ellipsoid acquires a dipole moment only because of the
applied external field.
We thus have to differentiate the expression (\ref{19}) with respect to $z$.
Performing the calculation along the same lines as above, we obtain
\begin{eqnarray} F_z&=&\frac{6p_zE_0}{B_{1-c^2/a^2}\left( \frac{3}{2},-\frac{1}{2}\right)}
\frac{(a^2-c^2)\sqrt{-\eta-c^2}}{(\xi+c^2)(\xi-\eta)} \nonumber\\
&&\quad\times \left[ 1-\frac{(\xi+a^2)(-\eta-c^2)}{(a^2-c^2)(\xi-\eta)}
+\frac23\frac{(\xi+c^2)(\eta+c^2)(\xi+\eta+2a^2)}{(a^2-c^2)(\xi-\eta)^2}
\right]. \label{22}
\end{eqnarray}
At $z=0$, the force vanishes as it should, since $\eta+c^2=0$ then.
Note that the force vanishes if $c/a\to0$, that is, for a disk, because
the integral representing the
incomplete beta function diverges in the limit.
(It is not to be interpreted as its analytic continuation.) This is not surprising,
for in the limit of a disk, the electric field is just $\mathbf{E}_0$, the
applied constant field. This is because inserting a perfectly conducting
sheet perpendicular to the field line has no effect on the boundary
conditions. See also the discussion in Chap.~4 of Ref.~\cite{radbook}.
As a small check, we consider the limit of a sphere, $c^2\to a^2$.
Then, according to Eq.~(\ref{16}), we have
\begin{equation}
B_{1-c^2/a^2}\left(\frac32,-\frac12\right)\to\frac23a^{-3}(a^2-c^2)^{3/2},
\end{equation}
and
\begin{equation}
\xi\approx \rho^2+z^2-c^2, \quad \eta=-c^2-\frac{\delta^2 z^2}{\rho^2+z^2},
\end{equation}
in terms of the ultimately vanishing quantity $\delta^2=a^2-c^2$.
Then we immediately obtain
\begin{equation}
F_z=3 p_z E_0\frac{a^3z}{(\rho^2+z^2)^{7/2}}(3\rho^2-2z^2).\label{dipsh}
\end{equation}
This result also follows immediately from the dipole-dipole interaction
energy
\begin{equation}
U=-\frac1{r^5}(3\mathbf{r\cdot p}_1\,\mathbf{r\cdot p}_2-r^2 \mathbf{p}_1\cdot
\mathbf{p}_2),
\end{equation}
when we take
\begin{equation} \mathbf{p}_1=p_z\mathbf{\hat z},\quad \mathbf{p}_2=a^3 E_0\mathbf{\hat z}.
\end{equation}
The force on the sphere (\ref{dipsh}) is attractive at large distance,
because the dipoles become essentially coaxial then, and repulsive at
small distance, because the case of parallel dipoles in a plane is approached
in that situation.
The same features hold for a general ellipsoid. For short distances,
$z^2\ll \rho^2-a^2+c^2$, we have
\begin{equation}
\xi=\rho^2-a^2+O(z^2),\quad \eta=-c^2-\frac{z^2(a^2-c^2)}{\rho^2-a^2+c^2}+
O(z^4),
\end{equation}
and then the force is repulsive,
\begin{equation}
z\to 0:\quad
F_z=\frac{6 p_z E_0}{B_{1-c^2/a^2}(3/2,-1/2)}\frac{z(a^2-c^2)^{3/2}}{(
\rho^2-a^2+c^2)^{5/2}},
\end{equation}
which reduces in the spherical case to
\begin{equation}
c\to a:\quad F_z=\frac{9p_zE_0 a^3 z}{\rho^5},
\end{equation}
which agrees with Eq.~(\ref{dipsh}). And in the large distance limit,
where $\xi\approx z^2$, $\eta\approx-a^2$, the force in general is attractive,
\begin{equation}
z\to\infty:\quad F_z=-\frac{4 p_z E_0(a^2-c^2)^{3/2}}{B_{1-c^2/a^2}(3/2,-1/2)}
\frac1{z^4},
\end{equation}
which again has the expected limit,
\begin{equation}
c\to a:\quad F_z=-\frac{6p_zE_0 a^3}{z^4}.
\end{equation}
\section{Interaction of anisotropic atom with anisotropic dielectric}
\label{sec6}
In view of the considerations of Sec.~\ref{sec2}, we might hope that
repulsion could be achieved if an anisotropic atom were placed above
an anisotropic dielectric medium. Consider such an atom, with polarizability
only in the $z$ direction, $\bm{\alpha}=\alpha\mathbf{\hat z\hat z}$,
a distance
$a$ above a dielectric with different permittivities in the $z$ direction and
the transverse directions,
\begin{equation}
\bm{\varepsilon}=\mbox{diag}(\varepsilon_\perp,\varepsilon_\perp,\varepsilon_\|).
\end{equation}
We will assume (see below) that $\varepsilon_\perp$, $\varepsilon_\|>1$.
The Casimir-Polder interaction is
\begin{equation}
E_{\rm CP}=-\alpha\int_{-\infty}^\infty d\zeta\left(\Gamma_{zz}-\Gamma^0_{zz}\right)
(\mathbf{R,R}),
\end{equation}
where the atom is located at $\mathbf{R}=(0,0,a)$.
Here we have subtracted the free-space contribution.
We can write the Green's dyadic in terms of a transverse Fourier transform,
\begin{equation}
\bm{\Gamma}(\mathbf{r,r'})=\int\frac{(d\mathbf{k}_\perp)}{(2\pi)^2}e^{i\mathbf{
k_\perp\cdot(r-r')_\perp}}\bm{\gamma}(z,z'),
\end{equation}
where (assuming that $\mathbf{k}_\perp$ lies in the $+x$
direction)
\begin{equation}
\bm{\gamma}(z,z')=\left(\begin{array}{ccc}
\frac1{\varepsilon_\perp}\frac\partial{\partial z}
\frac1{\varepsilon_\perp'}\frac\partial{\partial z'} g^H&0&
\frac{ik_\perp}{\varepsilon_\perp\varepsilon_\|'}\frac\partial{\partial z}g^H\\
0&-\zeta^2g^E&0\\
-\frac{ik_\perp}{\varepsilon'_\perp\varepsilon_\|}\frac\partial{\partial z'}g^H&
0&\frac{k_\perp^2}{\varepsilon_\|\varepsilon_\|'} g^H\end{array}\right).
\end{equation}
We have followed Ref.~\cite{Schwinger:1977pa}
and used the notation $\varepsilon=\varepsilon(z)$, $\varepsilon'=
\varepsilon(z')$. Here we have omitted $\delta$-function terms that
do not contribute in the point-splitting limit. The transverse electric
and transverse magnetic Green's functions satisfy the differential
equations
\begin{subequations}
\begin{eqnarray}
\left(-\frac{\partial^2}{\partial z^2}+k_\perp^2-\omega^2\varepsilon_\perp
\right)
g^E(z,z')=\delta(z-z'),\\
\left(-\frac{\partial}{\partial z}\frac1{\varepsilon_\perp}
\frac\partial{\partial z}
+\frac{k_\perp^2}{\varepsilon_\|}-\omega^2\right)
g^H(z,z')=\delta(z-z').
\end{eqnarray}
\end{subequations}
It is rather straightforward to solve these equations and find the Casimir-Polder
energy:
\begin{equation}
E_{\rm CP}=\frac{\alpha}{4\pi^2}\int_{-\infty}^\infty d\zeta\int (d\mathbf{k_\perp})
\frac{k_\perp^2}{2\kappa}\frac{\bar\kappa-\kappa}{\bar\kappa+\kappa}e^{-2\kappa a},
\label{aniso-atom-diel}
\end{equation}
where $\kappa^2=k_\perp^2-\omega^2$, $\bar\kappa=
\sqrt{(k_\perp^2-\omega^2\varepsilon_\|)/\varepsilon_\perp\varepsilon_\|}$.
Checks of this result are the following:
\begin{equation}
\varepsilon_\perp\to\infty: \quad E_{\rm CP}\to -\frac\alpha{8\pi a^4},
\end{equation}
one-third of the usual Casimir-Polder interaction of an isotropic atom with a perfect
conducting plate. This is what we would have for such an anisotropic atom above
a isotropic conducting plate, because taking $\varepsilon_\perp\to \infty$ imposes
the usual boundary condition that the tangential components of $\mathbf{E}$ vanish
on the surface.
In the other limit, we have no such simple correspondence,
\begin{equation}
\varepsilon_\|\to\infty: \quad E_{\rm CP}\to \frac\alpha{8\pi a^4}
\left(1+\frac32\sqrt{\varepsilon_\perp}-3\varepsilon_\perp+3\sqrt{\varepsilon_\perp}(\varepsilon_\perp-1)
\ln\frac{\sqrt{\varepsilon_\perp}+1}{\sqrt{\varepsilon_\perp}}\right),
\end{equation}
where the quantity in parentheses varies between $-1/2$ for $\varepsilon_\perp=1$
and $-1$ as $\varepsilon_\perp\to \infty$.
We can check that in all cases, if we ignore
dispersion, Eq.~(\ref{aniso-atom-diel})
yields an attractive result: $E_{\rm CP}$ scales like $a^{-4}$ times
a numerical integral which is always negative because $\bar\kappa^2-\kappa^2<0$.
Repulsion does not occur in this case because there is no breaking
of translational invariance in the transverse direction.
In fact,
the electromagnetic force density in an anisotropic nonmagnetic medium is
(see Ref.~\cite{brevikrev}, Eq.~(1.2a))
\begin{equation}
{\bf f}=-\frac1{8\pi}E_iE_k\nabla \varepsilon_{ik}.
\end{equation}
Assume that the single air-medium interface is flat, lying in the $xy$ plane.
Then the only nonvanishing component of the gradient $\nabla \varepsilon_{ik}$
is the vertical component $\partial_z \varepsilon_{ik}$. If the principal coordinate axes
for $\varepsilon_{ij}$ coincide with the $x$, $y$, $z$ axes,
then the surface force density $\int f_z \, dz$ (which is subsequently to be
integrated across the surface $z=0$), is directed upwards, because
$\varepsilon_{\parallel, \perp} >1$. The surface force acts in the direction of the optically thinner medium.
Now, momentum conservation of the total system asserts that the force on a dipole above the surface acts
in the downward direction. The dipole force has to be attractive.
That $\varepsilon >1$ for an isotropic medium is a thermodynamical result.
For an anisotropic medium, oriented such that the coordinate axes fall together
with the crystallographic axes, one must analogously have $\varepsilon_{\parallel, \perp}>1$.
See, for instance, Sec.~14 in Ref.~\cite{landau-lifshitz}.
Note the contrast with the force on a dipole outside a dielectric wedge, studied in Ref.~\cite{Milton:2011ni}.
In the latter case, the normal surface force on the inclined (lower) surface necessarily has a vertical
($z$) component that is downward directed. Momentum conservation for the total system thus no longer
forbids the force on the dipole to be repulsive.
\section{Conclusions}
Earlier, we observed that Casimir-Polder repulsion along a direction
perpendicular to the
symmetry axis of a semi-infinite planar conductor or a conducting wedge
and an anisotropically polarizable atom could be achieved in the region
close to the conductor \cite{Milton:2011ni}. Here we have shown that anisotropically
polarizable atoms can also repel in this sense, provided they are sufficiently
anisotropic, and have perpendicular principal axes. We further show that
such an
atom may be repelled by a conducting cylinder, provided,
at closest approach, it is sufficiently far
away from the cylinder, whereas no such phenomenon occurs for a sphere and an
anisotropic atom. We further discuss a new example of classical
repulsion by considering a polarized ellipsoid interacting with a dipole.
On the other hand, a system of an anisotropically polarizable atom interacting
via fluctuation forces with an anisotropic dielectric half-space
does not exhibit repulsion. Apparently, spatial anisotropy is also required
for repulsion between electric bodies.
\acknowledgments
We thank the US Department of Energy, and the US National Science Foundation,
for partial support of this research. The support of the ESF Casimir Network
is also acknowledged.
P.P. acknowledges the hospitality of the
University of Zaragoza. We thank E. K. Abalo for collaborative discussions.
We also thank K. V. Shajesh and M. Schaden for useful correspondence.
|
\section{Introduction}
This Letter reports on the search for diphoton ($\gamma\gamma$) events
with large missing transverse momentum (\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace) in \unit[1.07]{\ifb}\xspace of
proton-proton ($pp$) collision data at $\sqrt{s}=\unit[7]{TeV}$
recorded with the ATLAS detector in the first half of 2011, extending
a prior study performed with \unit[36]{\mbox{pb$^{-1}$}}~\cite{Aad:2011kz}. The
results are interpreted in the context of three models of new physics:
a general model of gauge-mediated supersymmetry breaking
(GGM)~\cite{Meade:2008wd,Buican:2008ws,Ruderman:2011vv}, a minimal
model of gauge-mediated supersymmetry breaking
(SPS8)~\cite{Allanach:2002nj}, and a model positing one universal
extra dimension
(UED)~\cite{Appelquist:2000nn,Macesanu:2002ew,Macesanu:2005jx}.
\section{Supersymmetry}
\label{sec:susy}
Supersymmetry~(SUSY)~\cite{Golfand:1971iw,Neveu:1971rx,Ramond:1971gb,Volkov:1973ix,Wess:1974tw}
introduces a symmetry between fermions and bosons, resulting in a SUSY
partner (sparticle) with identical quantum numbers except a difference
by half a unit of spin for each Standard Model (SM) particle. As none
of these sparticles have been observed, SUSY must be a broken symmetry
if realised in nature. Assuming $R$-parity
conservation~\cite{Fayet:1977yc,Farrar:1978xj}, sparticles have to be
produced in pairs. These would then decay through cascades involving
other sparticles until the lightest SUSY particle (LSP) is produced,
which is stable.
In gauge-mediated SUSY breaking~(GMSB)
models~\cite{Dine:1981,Dimopoulus:1981,AlvarezGaume:1981wy,Nappi:1982,Dine:1995ag}
the LSP is the gravitino \ensuremath{\tilde{G}}\xspace. GMSB experimental signatures are
largely determined by the nature of the next-to-lightest SUSY
particle~(NLSP), which for a large part of the GMSB parameter space is
the lightest neutralino \ensuremath{\tilde{\chi}^{0}_{1}}\xspace. Should the lightest neutralino
have similar couplings as the SM U(1) gauge boson, also referred to as
``bino'' in this case, the final decay in the cascade would
predominantly be $\ensuremath{\tilde{\chi}^{0}_{1}}\xspace\to\gamma\ensuremath{\tilde{G}}\xspace$, with two cascades
per event, leading to final states with $\gamma\gamma+\MET$, where
$\MET$ results from the undetected gravitinos.
Searches for GMSB performed at the Tevatron~\cite{Aaltonen:2009tp,Abazov:2010us}
were optimized to be sensitive to a minimal GMSB model
(SPS8)~\cite{Allanach:2002nj}. To reduce the number of free parameters
in this model, several assumptions are made. These assumptions lead to
a mass hierarchy in which squarks and gluinos are much heavier than
the lightest neutralino and chargino \ensuremath{\tilde{\chi}^{\pm}_{1}}\xspace. The SUSY breaking mass
scale felt by the low-energy sector, $\Lambda$, is the only free
parameter of the SPS8 model.
The other model parameters are fixed to the following values: the
messenger mass $M_{\mathrm{mess}}=2\Lambda$, the number of copies of
$5+\bar{5}$ SU(5) messengers $\ensuremath{N_{5}}\xspace=1$, the ratio of the
vacuum expectation values of the two Higgs doublets $\tan\beta=15$, and the Higgs sector mixing
parameter $\mu>0$. The NLSP is assumed to decay promptly
($c\tau_{\mathrm{NLSP}} < \unit[0.1]{mm}$).
At the present LHC energy the main contribution to the production
cross section in the SPS8 model is via gaugino pair production, i.e.
production of \ensuremath{\tilde{\chi}^{0}_{2}}\xspace\ensuremath{\tilde{\chi}^{\pm}_{1}}\xspace or \ensuremath{\tilde{\chi}^{0}_{2}}\xspace\neutralinotwo
pairs. The contribution from gluino and/or squark pairs is below
\unit[10]{\%} of the production cross section due to their high
masses.
Besides the two photons and the two gravitinos, jets, leptons, and gauge
bosons may be produced in the cascades.
This Letter presents the first limits on the SPS8 model at the LHC.
Furthermore, a GGM SUSY model is considered in which the gluino and
neutralino masses are treated as free parameters. The other sparticle
masses are fixed at $\sim\unit[1.5]{TeV}$, leading to a dominant
production mode at $\sqrt{s} = \unit[7]{TeV}$ of a pair of gluinos via
the strong interaction that would decay via cascades into the
bino-like neutralino NLSP. Jets may be produced in the cascades from
the gluino decays if kinematically allowed. Further model parameters
are fixed to $\tan\beta=2$ and $c\tau_{\mathrm{NLSP}} <
\unit[0.1]{mm}$.
The decay into the wino-like neutralino NLSP is possible and was
studied by the CMS Collaboration~\cite{Chatrchyan:2011ah}.
\section{Extra dimensions}
UED models postulate the existence of additional spatial dimensions in
which all SM particles can propagate, leading to the existence of a
series of excitations for each SM particle, known as a Kaluza-Klein
(KK) tower. This analysis considers the case of a single UED, with
compactification radius (size of the extra dimension) $R \approx
\unit[1]{TeV^{-1}}$.
At the LHC, the main UED process would be the production via the
strong interaction of a pair of first-level KK quarks and/or
gluons~\cite{Macesanu:2002db}. These would decay via cascades
involving other KK particles until reaching the lightest KK particle
(LKP), i.e. the first level KK photon $\gamma^*$. SM particles as
quarks, gluons, leptons, and gauge bosons may be produced in the
cascades. If the UED model is embedded in a larger space with $N$
additional eV$^{-1}$-sized dimensions accessible only to
gravity~\cite{DeRujula:2000he}, with a $(4 + N)$-dimensional Planck
scale ($M_D$) of a few \unit{TeV}, the LKP would decay gravitationally
via $\gamma ^* \rightarrow \gamma + G$. $G$ represents a tower of
eV-spaced graviton states, leading to a graviton mass between $0$ and
$1/R$.
With two decay chains per event, the final state would contain $\gamma
\gamma + \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace$, where \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace results from the escaping gravitons.
Up to $1/R \sim\unit[1]{TeV}$, the branching ratio to the diphoton and
\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace final state is close to \unit[100]{\%}. As $1/R$ increases, the
gravitational decay widths become more important for all KK particles
and the branching ratio into photons decreases, e.g. to \unit[50]{\%}
for $1/R=\unit[1.5]{TeV}$~\cite{Macesanu:2002ew}.
The UED model considered here is defined by specifying $R$ and
$\Lambda$, the ultraviolet cut-off used in the calculation of
radiative corrections to the KK masses. This analysis sets $\Lambda $
such that $\Lambda R = 20$. The $\gamma^*$ mass is insensitive to
$\Lambda$, while other KK masses typically change by a few per cent
when varying $\Lambda R$ in the range $10-30$. For $1/R =
\unit[1200]{GeV}$, the masses of the first-level KK photon, quark, and
gluon are $1200$, $1387$ and \unit[1468]{GeV},
respectively~\cite{ElKacimi:2009zj}.
Further details of the model are given in Ref.\xspace~\cite{Aad:2011kz}.
\section{Simulated samples}
For the GGM model, the SUSY mass spectra were calculated using
{\tt SUSPECT}\xspace 2.41~\cite{Djouadi:2002ze} and
{\tt SDECAY}\xspace~1.3~\cite{Muhlleitner:2003vg}. The Monte Carlo~(MC) signal
samples were produced using {\tt PYTHIA}\xspace~6.423~\cite{pythia} with {\tt
MRST2007 LO${}^{*}$}~\cite{Sherstnev:2007nd} parton distribution
functions~(PDF).
Cross sections were calculated at next-to-leading order (NLO) using
{\tt PROSPINO}\xspace 2.1~\cite{prospino2,Beenakker:1996ch}.
For the SPS8 model, the SUSY mass spectra were calculated using
{\tt ISAJET}\xspace~7.80~\cite{isajet}. The MC signal samples were produced using
{\tt HERWIG++}\xspace~2.4.2~\cite{Bahr:2008pv} with {\tt MRST2007 LO${}^{*}$} PDF.
NLO cross sections were calculated using {\tt PROSPINO}\xspace.
In the case of the UED model, MC signal samples were generated using
the UED model as implemented at leading order (LO) in
{\tt PYTHIA}\xspace~\cite{ElKacimi:2009zj}.
The ``irreducible'' background from $(W\to\ell\nu)\gamma\gamma$ and
$(Z\rightarrow\nu\nu)\gamma\gamma$ production was simulated at LO
using {\tt MadGraph}\xspace~4~\cite{Alwall:2007st} with {\tt
CTEQ6L1}~\cite{cteq6m} PDF. Parton showering and fragmentation were
simulated with {\tt PYTHIA}\xspace. NLO cross sections and scale uncertainties
from Ref.\xspace~\cite{Bozzi:2011en,Bozzi:2011wwa} were used.
In all cases the underlying event was simulated within the respective
generator.
All samples were processed through the {\tt GEANT4}-based
simulation~\cite{geant4} of the ATLAS detector~\cite{Aad:2010wq}. In
addition, the signal samples were overlaid with simulated minimum bias
events to model the average number of six $pp$ interactions per bunch
crossing (pile-up) experienced during the considered data-taking
period. More details may be found in Ref.\xspace~\cite{Aad:2011kz}.
\section{ATLAS detector}
The ATLAS detector~\cite{Aad:2008zzm} is a multi-purpose apparatus
with a forward-backward symmetric cylindrical geometry and nearly
4$\pi$ solid angle coverage. Closest to the beamline are tracking
devices comprised of layers of silicon-based pixel and strip detectors
covering $\left|\eta\right|<2.5$\footnote{ATLAS uses a right-handed
coordinate system with its origin at the nominal interaction point
(IP) in the centre of the detector and the $z$-axis along the beam
pipe. The $x$-axis points from the IP to the centre of the LHC ring,
and the $y$-axis points upward. Cylindrical coordinates $(R,\phi)$
are used in the transverse plane, $\phi$ being the azimuthal angle
around the beam pipe. The pseudorapidity is defined in terms of the
polar angle $\theta$ as $\eta=-\ln\tan(\theta/2)$.} and straw-tube
detectors covering $\left|\eta\right|<2.0$, located inside a thin
superconducting solenoid that provides a \unit[2]{T} magnetic field.
The straw-tube detectors also provide discrimination between electrons
and charged hadrons based on transition radiation. Outside the
solenoid, fine-granularity lead/liquid-argon (LAr) electromagnetic
(EM) calorimeters provide coverage for $\left|\eta\right| < 3.2$ to
measure the energy and position of electrons and photons. In the
region $\left|\eta\right| < 2.5$, the EM calorimeters are segmented
into three layers in depth. The second layer, in which most of the EM
shower energy is deposited, is divided into cells of granularity of
$\Delta \eta \times \Delta \phi = 0.025 \times 0.025$. The first layer
is segmented with finer granularity to provide discrimination between
single photons and overlapping photons coming from the decays of
neutral mesons. A presampler, covering $\left|\eta\right| < 1.8$, is
used to correct for energy lost upstream of the EM calorimeter. An
iron/scintillating-tile hadronic calorimeter covers the region $|\eta|
< 1.7$, while copper and liquid-argon technology is used for hadronic
calorimeters in the end-cap region $1.5 < |\eta| < 3.2$. In the
forward region $3.2 < |\eta| < 4.5$ liquid-argon calorimeters with
copper and tungsten absorbers measure the electromagnetic and hadronic
energy. A muon spectrometer consisting of three superconducting
toroidal magnet systems, tracking chambers, and detectors for
triggering surrounds the calorimeter system.
\section{Object reconstruction}
The reconstruction of converted and unconverted photons and of
electrons is described in Refs.~\cite{Aad:2010sp} and
\cite{Aad:2011mk}, respectively.
Converted photons have EM calorimeter clusters matched to tracks
coming from a conversion vertex. A conversion vertex is either a
vertex that has two tracks with large transition radiation in the
straw-tube detector and an invariant mass of the two tracks consistent
with a massless particle, i.e. a photon, or one track with large
transition radiation that has no associated hits in the pixel layer
closest to the beam line. Electrons have a track matched to the EM
calorimeter cluster, and the track must have hits in the silicon
detectors, momentum not smaller than one tenth the cluster energy, and
transverse momentum of at least $\unit[2]{GeV}$. Clusters matched to
neither a track or tracks coming from a conversion vertex nor an
electron track as described above are classified as unconverted
photons. A heuristic using the pixel hits closest to the beam line and
the track momenta is applied to choose between the photon and electron
interpretation in cases where the object can be both.
Photon candidates were required to be within $\left|\eta\right| <
1.81$, the value being chosen by an optimisation of the signal
acceptance versus background rejection, and to be outside the
transition region $1.37 < \left|\eta\right| < 1.52$ between the barrel
and the end-cap calorimeters. The analysis used ``loose'' and
``tight'' photon selections~\cite{Aad:2010sp}. The loose photon
selection includes a limit on the fraction of the energy deposit in
the hadronic calorimeter as well as a requirement that the transverse
width of the shower, measured in the middle layer of the EM
calorimeter, be consistent with the narrow shape expected for an EM
shower. The tight photon selection additionally uses shape
information from the first layer to distinguish between isolated
photons and photons from the decay of neutral mesons.
The reconstruction of \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace is based on energy deposits in calorimeter
cells inside three-dimensional clusters with $|\eta | < 4.5$ and is
corrected for contributions from muons, if any~\cite{Aad:2011re}. The
cluster energy is calibrated to correct for the non-compensating
calorimeter response, energy loss in dead material, and out-of-cluster
energy.
Jets were reconstructed using the anti-$k_t$ jet
algorithm~\cite{Cacciari:2008gp} with four-momentum recombination and
radius parameter $R=0.4$ in $\eta$-$\phi$ space. They were required to
have $\ensuremath{p_{\mathrm{T}}}\xspace > \unit[25]{GeV}$ and $|\eta| < 2.8$.
\section{Data analysis}
The data sample, corresponding to an integrated luminosity of
\unit[$(1.07\pm 0.04)$]{\ifb}\xspace, was selected by a trigger requiring two loose photon
candidates with a transverse energy ($\ensuremath{E_{\mathrm{T}}}\xspace$) above \unit[20]{GeV}.
In the offline analysis events were retained if they contained at
least two tight photon candidates with $\ensuremath{E_{\mathrm{T}}}\xspace > \unit[25]{GeV}$. In
addition, a photon isolation cut was applied, whereby the \ensuremath{E_{\mathrm{T}}}\xspace deposit
in a cone of radius $0.2$ in the $\eta$-$\phi$ space around the centre
of the cluster, excluding the cells belonging to the cluster, had to
be less than \unit[5]{GeV}. The \ensuremath{E_{\mathrm{T}}}\xspace was corrected for leakage from the
photon energy outside the cluster and for soft energy deposits from
pile-up interactions.
A cut of $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace>\unit[125]{GeV}$~\cite{Aad:2011kz} defined the signal
region. Preference was given to a common signal region for the three
models considered.
A total of 27293 $\gamma\gamma$ candidate events were observed passing
all selections except the \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace cut. The \ensuremath{E_{\mathrm{T}}}\xspace distribution of the
leading photon for events in this sample is shown in
Fig.\xspace~\ref{fig:PhoPt}. Also shown are the \ensuremath{E_{\mathrm{T}}}\xspace spectra obtained from GGM
MC samples for $\mass{\ensuremath{\tilde{g}}\xspace}=\unit[800]{GeV}$ and
$\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace}=\unit[400]{GeV}$, from SPS8 MC samples with
$\Lambda = \unit[140]{TeV}$, and from UED MC samples for $1/R =
\unit[1200]{GeV}$, representing model parameters near the expected
exclusion limit. After the $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace > \unit[125]{GeV}$ cut, 5\xspace
candidate events survived.
\begin{figure}[tb]
\begin{center}
\includegraphics[width=0.48\textwidth]{ph_mX400_mg800_R1200_pt1.eps}
\end{center}
\caption{The \ensuremath{E_{\mathrm{T}}}\xspace spectrum of the leading photon in the $\gamma\gamma$
candidate events in the data (points, statistical uncertainty only) together with the spectra from simulated GGM
($\mass{\ensuremath{\tilde{g}}\xspace},\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace} = \unit[(800,400)]{GeV}$), SPS8
($\Lambda = \unit[140]{TeV}$), and UED
($1/R = \unit[1200]{GeV}$) samples, prior to the application of the
$\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace>\unit[125]{GeV}$ cut. The signal samples are scaled by a factor of 100 for clarity.
\label{fig:PhoPt}
}
\end{figure}
\begin{figure}[tb]
\begin{center}
\includegraphics[width=0.48\textwidth]{DiPhotonMetPlot.eps}
\end{center}
\caption{\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace spectra for the $\gamma\gamma$ candidate events in data (points,
statistical uncertainty only) and the estimated QCD background
(normalised to the number of $\gamma\gamma$ candidates with $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace <
\unit[20]{GeV}$), the $\ensuremath{W(\to e\nu)+\mathrm{jets}/\gamma}\xspace$ and $\ensuremath{\ttbar(\to e\nu)+\mathrm{jets}}\xspace$ backgrounds as estimated from
the electron-photon control sample, and the irreducible
background of \ensuremath{Z(\to \nu\bar{\nu})+\gamma\gamma}\xspace and \ensuremath{W(\to \ell\nu)+\gamma\gamma}\xspace. Also shown are the expected signals
from GGM ($\mass{\ensuremath{\tilde{g}}\xspace},\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace} =
\unit[(800,400)]{GeV}$), SPS8 ($\Lambda = \unit[140]{TeV}$), and UED
($1/R = \unit[1200]{GeV}$) samples.
\label{fig:bkgnd}
}
\end{figure}
\section{Background estimation}
Following the procedure described in Ref.\xspace~\cite{Aad:2011kz}, the
contribution to large~\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace diphoton events from SM sources can be
grouped into two primary components and estimated with dedicated
control samples using data. The first of these components, referred
to as ``QCD background'' for brevity, arises from a mixture of
processes that include $\gamma \gamma$ production as well as $\gamma $
+ jet and multijet events with at least one jet mis-reconstructed as a
photon. The second background component is due to $W+X$ and \antibar{t}\xspace
events, where mis-reconstructed photons can arise from electrons and jets, for
which final-state neutrinos produce significant \MET.
In order to estimate the QCD background from $\gamma\gamma$, $\gamma$
+ jet, and multijet events, a ``\ensuremath{\mathrm{QCD}}\xspace control sample'' was extracted
from the diphoton trigger sample by selecting events for which at
least one of the photon candidates does not pass the tight photon
identification. Electrons were vetoed to remove contamination from
$W\to e\nu$ decays.
The QCD background contamination in the signal region $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace >
\unit[125]{GeV}$ was obtained from this QCD template after normalising
it to data in the region $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace < \unit[20]{GeV}$. This gives a QCD
background expectation in the signal region of
$0.8\pm0.3(\mathrm{stat})$ events.
An alternate model for the QCD background was
obtained using a sample of dielectron events, with no jets, selected
by requiring two electrons with $\ensuremath{E_{\mathrm{T}}}\xspace>\unit[25]{GeV}$ and
$\left|\eta\right| < 1.81$ and an invariant mass consistent with the
$Z$ boson mass.
As confirmed by MC simulation, the \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace spectrum of this $\ensuremath{Z \rightarrow ee}\xspace$ sample
with no additional jets, which is dominated by the calorimeter
response to two genuine EM objects, accurately represents the \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace
spectrum of SM $\gamma\gamma$ events. This spectrum was normalised in
the same way as the QCD control sample.
A systematic uncertainty of 0.6 events was assigned as the systematic
uncertainty on the background prediction from the relative fractions
of $\gamma\gamma$, $\gamma$ + jet, and multijet events using the
difference between the background estimates obtained using the QCD and
the \ensuremath{Z \rightarrow ee}\xspace templates, yielding the result of
$0.8\pm0.3(\mathrm{stat})\pm0.6(\mathrm{syst})$ events.
The \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace spectra of the QCD background and the $\gamma\gamma$ sample
are shown in Fig.\xspace~\ref{fig:bkgnd}.
\begin{figure}[bt]
\begin{center}
\includegraphics[width=0.48\textwidth]{ElectronPhotonMetPlot.eps}
\end{center}
\caption{\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace spectrum for the electron-photon control sample in data (points, statistical uncertainty only), normalised according to the probability for an
electron to be mis-reconstructed as a tight photon, compared to the expected
backgrounds displayed by components (stacked histograms).
For the purpose of this comparison, the
expected contributions from $\ensuremath{W(\to e\nu)+\mathrm{jets}/\gamma}\xspace$ and \ensuremath{\ttbar(\to e\nu)+\mathrm{jets}}\xspace events are taken from
MC simulation.}
\label{fig:ElPho}
\end{figure}
\begin{table*}[t!]
\caption{Number of observed $\gamma\gamma$ candidates in various \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace
ranges in the data, as well as the expected numbers of SM background
events estimated from the QCD and electron-photon control samples
and, for the irreducible \ensuremath{Z(\to \nu\bar{\nu})+\gamma\gamma}\xspace and \ensuremath{W(\to \ell\nu)+\gamma\gamma}\xspace processes, from MC
simulation. Also shown are the expected numbers of signal events
from GGM with $(\mass{\ensuremath{\tilde{g}}\xspace},\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace}) =
\unit[(800,400)]{GeV}$, SPS8 with $\Lambda = \unit[140]{TeV}$, and
UED with $1/R = \unit[1200]{GeV}$. The uncertainties are
statistical only. The $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace < \unit[20]{GeV}$ region (first row) is
used to normalise the QCD background to the number of observed
$\gamma \gamma$ candidates.
}
\label{tab:events}
\center
\begin{tabular}{c|r|cccc|ccc}
\hline
\hspace{-0.06in}\MET range & \multicolumn{1}{c|}{Data} & \multicolumn{4}{c|}{Predicted background events} & \multicolumn{3}{c}{Expected signal events} \\
$[$GeV$]$ & events & Total & QCD & \makebox[57pt]{$W/\antibar{t}\xspace(\rightarrow e \nu)+X$} & Irreducible & GGM & SPS8 & UED \\ \hline
\hspace{0.07in}0 - 20 & 20881 & \hspace{0.13in}- & \hspace{0.01in}- & \hspace{0.06in}- & - & $\hspace{0.06in}0.20\pm 0.05$ & $\hspace{0.06in}0.22\pm 0.04$ & $0.02\pm 0.01$ \\ \hline
\hspace{0.00in}20 - 50 & 6304 & $\hspace{0.00in}5968\pm 29$ & $\hspace{-0.12in}5951\pm28$ & $\hspace{0.00in}13.3\pm 8.1$ & $3.55\pm 0.35$ & $\hspace{0.06in}0.45 \pm 0.08$ & $\hspace{0.06in}1.53\pm 0.10$ & $0.11\pm 0.01$ \\
\hspace{0.00in}50 - 75 & 86 & $\hspace{0.06in}87.1\pm 3.3$ & $\hspace{-0.06in}60.9\pm2.8$ & $\hspace{0.00in}25.2\pm 1.7$ & $1.01\pm 0.16$ & $\hspace{0.06in}0.48 \pm 0.08$ & $\hspace{0.06in}2.19\pm 0.12$ & $0.14\pm 0.01$ \\
\hspace{0.06in}75 - 100 & 11 & $\hspace{0.06in}14.7\pm 1.2$ & $\hspace{0.0in}6.7\pm0.9$ & $\hspace{0.06in}7.4\pm 0.8$ & $0.52\pm 0.10$ & $\hspace{0.06in}0.75 \pm 0.10$ & $\hspace{0.06in}2.09\pm 0.11$ & $0.15\pm 0.01$ \\
\hspace{0.00in}100 - 125 & 6 & $\hspace{0.12in}4.9\pm 0.7$ & $\hspace{0.0in}1.6\pm0.4$ & $\hspace{0.06in}3.0\pm 0.5$ & $0.32\pm 0.08$ & $\hspace{0.06in}1.20 \pm 0.12$ & $\hspace{0.06in}2.53\pm 0.13$ & $0.29\pm 0.02$ \\
\hspace{0.22in}$>$ 125 & 5 & $\hspace{0.12in}4.1\pm 0.6$ & $\hspace{0.0in}0.8\pm 0.3$ & $\hspace{0.06in}3.1\pm 0.5$ & $0.23\pm 0.05$ & $17.2 \pm 0.5$ & $12.98\pm 0.28$ & $9.67\pm 0.11$ \\
\hline
\end{tabular}
\end{table*}
The second significant background contribution, from $W+X$ and \antibar{t}\xspace
events, was estimated via an ``electron-photon'' control sample
composed of events with at least one photon and one electron, each with
$\ensuremath{E_{\mathrm{T}}}\xspace>\unit[25]{GeV}$, and scaled by the probability for an electron to
be mis-reconstructed as a tight photon, as estimated from a study of the
$Z$ boson in the $ee$ and $e\gamma$ sample. The scaling factor varies
between \unit[5]{\%} and \unit[17]{\%} as a function of $\eta$, since
it depends on the amount of material in front of the calorimeter.
Events with two or more photons were vetoed from the control sample to keep it
orthogonal to the signal sample. In case of more than one electron,
the one with the highest \ensuremath{p_{\mathrm{T}}}\xspace was used.
The \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace spectrum for the scaled electron-photon control sample is
shown in Fig.\xspace~\ref{fig:ElPho}, where it is compared to the expected
contributions from various background sources as computed from MC
simulation. The electron-photon control sample has a significant
contamination from \ensuremath{Z \rightarrow ee}\xspace events, in which one electron is mis-reconstructed as a photon,
and from QCD processes mentioned above. Both of these contaminations
must be subtracted in order to extract the contribution to the \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace
distribution from events with genuine \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace, such as $W+X$ and \antibar{t}\xspace.
The contribution from QCD and \ensuremath{Z \rightarrow ee}\xspace events was estimated by normalising
the \ensuremath{\mathrm{QCD}}\xspace control sample to the scaled electron-photon \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace
distribution in the region $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace < \unit[20]{GeV}$ where they
dominate, as shown in Fig.\xspace~\ref{fig:ElPho}. This distribution was
then subtracted from the scaled electron-photon control sample,
yielding a prediction for the contribution to the high-\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace diphoton
sample from $W+X$ and \antibar{t}\xspace events. This procedure led to an
estimate of the background from $W+X$ and \antibar{t}\xspace production of $3.1
\pm 0.5(\mathrm{stat})$ events in the signal region. A systematic
uncertainty of 0.06 events was assigned by using the \ensuremath{Z \rightarrow ee}\xspace template in
place of the \ensuremath{\mathrm{QCD}}\xspace template when subtracting the contamination due to
\ensuremath{Z \rightarrow ee}\xspace and QCD processes.
The contribution from $WW$ events to the electron-photon control
sample was estimated using MC simulation and found to be negligible.
A parallel study using MC samples of \ensuremath{W(\to e\nu)+\mathrm{jets}/\gamma}\xspace and \ensuremath{\ttbar(\to e\nu)+\mathrm{jets}}\xspace, rather than
the electron-photon control sample, gave an estimate of $1.8 \pm
1.2(\mathrm{stat})$ background events. The difference was taken as an
estimate of the systematic uncertainty, yielding the result of
$3.1\pm0.5(\mathrm{stat})\pm1.4(\mathrm{syst})$ events. Also included
in the quoted systematic uncertainty is the relative uncertainty ($\pm
\unit[10]{\%}$) on the probability for an electron to be mis-reconstructed as a photon.
A small irreducible background of $0.23 \pm 0.05 (\mathrm{stat}) \pm
0.04 (\mathrm{syst})$ events from \ensuremath{Z(\to \nu\bar{\nu})+\gamma\gamma}\xspace and \ensuremath{W(\to \ell\nu)+\gamma\gamma}\xspace events was estimated
from MC simulation. The systematic uncertainty accounts for variations
in the factorisation and renormalisation scales in the NLO
calculations.
The contamination from cosmic-ray muons was found to be negligible.
Figure~\ref{fig:bkgnd} shows the \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace spectrum of the selected $\gamma
\gamma$ candidates, superimposed on the estimated backgrounds.
Table~\ref{tab:events} summarises the number of observed $\gamma
\gamma$ candidates, the expected backgrounds, and three representative
GGM, SPS8, and UED signal expectations, in several \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace ranges. No
indication of an excess at high \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace values, where the signal is
expected to dominate, is observed.
\section{Signal efficiencies and systematic uncertainties}
The GGM signal efficiency was determined using MC simulation over an
area of the GGM parameter space that ranges from \unit[400]{GeV} to
\unit[1200]{GeV} for the gluino mass, and from $\unit[50]{GeV}$ to
within $\unit[20]{GeV}$ of the gluino mass for the neutralino mass.
The efficiency increases smoothly from $\unit[5.5]{\%}$ to
$\unit[31]{\%}$ for $(\mass{\ensuremath{\tilde{g}}\xspace},\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace}) =
\unit[(400,50)]{GeV}$ to $\unit[(1200,1100)]{GeV}$.
The SPS8 signal efficiency increases smoothly from \unit[9.2]{\%}
($\Lambda = \unit[80]{TeV}$) to \unit[29.4]{\%} ($\Lambda =
\unit[220]{TeV}$).
The UED signal efficiency, also determined using MC simulation,
increases smoothly from \unit[48.9]{\%} ($1/R = \unit[1000]{GeV}$) to
\unit[52.6]{\%} ($1/R = \unit[1500]{GeV}$).
\begin{table}[b!]
\caption{Relative systematic uncertainties on the expected signal
yield for GGM with $(\mass{\ensuremath{\tilde{g}}\xspace},\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace}) =
\unit[(800,400)]{GeV}$, SPS8 with $\Lambda=\unit[140]{TeV}$, and UED with $1/R = \unit[1200]{GeV}$. No
PDF and scale uncertainties are given for the UED case as the cross
section is evaluated to LO. }
\label{tab:systematics}
\center
\begin{tabular}{lrrr}
\hline
Source of uncertainty & \multicolumn{3}{c}{Uncertainty} \\
& \multicolumn{1}{c}{GGM} & \multicolumn{1}{c}{SPS8} & \multicolumn{1}{c}{UED} \\ \hline
Integrated luminosity & $\hspace{0.04in}3.7$\% & $\hspace{0.04in}3.7$\% & $\hspace{0.04in}3.7$\% \\
Trigger & $\hspace{0.04in}0.6$\% & $\hspace{0.04in}0.6$\% & $\hspace{0.04in}0.6$\% \\
Photon identification & $\hspace{0.02in}3.9$\% & $\hspace{0.02in}3.9$\% & $\hspace{0.02in}3.7$\% \\
Photon isolation & $\hspace{0.02in}0.6$\% & $\hspace{0.02in}0.6$\% & $\hspace{0.02in}0.5$\% \\
Pile-up & $\hspace{0.02in}1.3$\% & $\hspace{0.02in}1.3$\% & $\hspace{0.02in}1.6$\% \\
$\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace$ reconstruction and scale & $\hspace{0.02in}1.7$\% & $\hspace{0.02in}5.6$\% & $\hspace{0.02in}0.7$\% \\
LAr readout & $\hspace{0.02in}1.0$\% & $\hspace{0.02in}0.7$\% & $\hspace{0.02in}0.4$\% \\
Signal MC statistics & $\hspace{0.02in}2.9$\% & $\hspace{0.02in}2.3$\% & $\hspace{0.02in}1.8$\% \\
\hline
Total signal uncertainty & $\hspace{0.02in}6.6$\% & $\hspace{0.02in}8.3$\%& $\hspace{0.02in}6.0$\% \\
\hline
PDF and scale & $\hspace{0.02in}31$\% & $\hspace{0.02in}5.5$\% & $\hspace{0.02in}-$ \\
\hline
Total & $\hspace{0.02in}32$\% & $\hspace{0.02in}10$\% & $\hspace{0.02in}6.0$\% \\
\hline
\end{tabular}
\end{table}
The various relative systematic uncertainties on the GGM, SPS8, and UED
signal cross sections are summarised in Table\xspace~\ref{tab:systematics} for
the chosen GGM, SPS8, and UED reference points. The uncertainty on the
luminosity is \unit[3.7]{\%}~\cite{lumi2011,Aad:2011dr}.
The trigger efficiency of the required diphoton trigger was estimated from
the efficiency of the corresponding single photon trigger, which was
estimated using a bootstrap method~\cite{Aad:2011xs}. The result
is $99.92^{+0.04}_{-0.18} \%$ for events passing all selections except
the final \ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace cut.
To estimate the systematic uncertainty due to the unknown composition
of the data sample, the trigger efficiency was also evaluated on MC
events using mis-reconstructed photons from filtered multijet samples and photons
from signal (SUSY and UED) samples. A conservative systematic
uncertainty of \unit[0.6]{\%} was derived from the difference between
the obtained efficiencies. Uncertainties on the photon selection, the
photon energy scale, and the detailed material composition of the
detector, as described in Ref.\xspace~\cite{Aad:2011kz}, result in an
uncertainty of \unit[3.9]{\%} for the GGM and SPS8 signals and
\unit[3.7]{\%} for the UED signal. The uncertainty from the photon
isolation was estimated by varying the energy leakage and the pile-up
corrections independently, resulting in an uncertainty of
\unit[0.6]{\%} for GGM and SPS8 and \unit[0.5]{\%} for UED. The
influence of pile-up on the signal efficiency, evaluated by comparing
GGM/SPS8 (UED) MC samples with different pile-up configurations, leads
to a systematic uncertainty of \unit[1.3]{\%}(\unit[1.6]{\%}).
Systematic uncertainties due to the $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace$ reconstruction, estimated
by varying the cluster energies within established ranges and the
$\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace$ resolution between the measured performance and MC
expectations, contribute an uncertainty of \unit[0.1]{\%} to
\unit[12.4]{\%}~(GGM), \unit[1.7]{\%} to \unit[13.8]{\%} (SPS8), and
\unit[0.5]{\%} to \unit[1.5]{\%}~(UED). A systematic uncertainty was
also assigned to account for temporary failures of the LAr calorimeter
readout during part of the data-taking period, which was not modelled
in the MC samples. Electrons and photons were removed from the
afflicted area, but jets, being larger objects, were not. Jet energy
corrections were therefore applied. Varying these corrections over
their range of uncertainty results in systematic uncertainties of
\unit[1.0]{\%}, \unit[0.7]{\%}, and \unit[0.4]{\%} for GGM, SPS8, and
UED, respectively.
Added in quadrature, the total systematic uncertainty on the signal
yield varies between \unit[6.3]{\%} and \unit[15]{\%}~(GGM),
\unit[6.2]{\%} and \unit[15]{\%}~(SPS8), and \unit[5.8]{\%} and
\unit[6.0]{\%} (UED).
The PDF uncertainties on the GGM (SPS8) cross sections were evaluated
by using the {\tt CTEQ6.6M} PDF error sets~\cite{Stump:2003yu} in the
{\tt PROSPINO}\xspace cross section calculation and range from \unit[12]{\%} to
\unit[44]{\%} (\unit[4.7]{\%} to \unit[6.6]{\%}). The factorisation
and renormalisation scales in the NLO {\tt PROSPINO}\xspace calculation were
increased and decreased by a factor of two, leading to a systematic
uncertainty between \unit[16]{\%} and \unit[23]{\%} (\unit[1.7]{\%}
and \unit[6.7]{\%}) on the expected cross sections.
The different impact of the PDF and scale uncertainties of the GGM and
SPS8 yields is related to the different production mechanisms in the
two models (see Section~\ref{sec:susy}).
In the case of UED, the PDF uncertainties were evaluated by using the
{\tt MSTW2008 LO}~\cite{MSTW2008} PDF error sets in the LO cross
section calculation and are about \unit[4]{\%}. The scale of
$\alpha_s$ in the LO cross section calculation was increased and
decreased by a factor of two, leading to a systematic uncertainty of
\unit[4.5]{\%} and \unit[9]{\%}, respectively. NLO calculations are
not yet available, but are expected to be much larger than the PDF and
scale uncertainties. Thus, the LO cross sections were used for the
limit calculation without any theoretical uncertainty, and the effect
of PDF and scale uncertainties on the final limit is given separately.
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.48\textwidth]{GGM_limit.eps}
\end{center}
\caption{Expected and observed \unit[95]{\%} CL\xspace lower limits on the
gluino mass as a function of the neutralino mass in the GGM model
with a bino-like lightest neutralino NLSP (the grey area
indicates the region where the NLSP is the gluino, which is not
considered here). The other sparticle masses are fixed to
$\sim\unit[1.5]{TeV}$. Further model parameters are $\tan\beta=2$
and $c\tau_{\mathrm{NLSP}} < \unit[0.1]{mm}$. The
previous ATLAS~\protect\cite{Aad:2011kz} and
CMS~\protect\cite{Chatrchyan:2011wc} limits are also shown.
\label{fig:ggm:limit}
}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.44\textwidth]{SPS8_LLRlimit_Cls_Prelim.eps}
\end{center}
\caption{Expected and observed \unit[95]{\%}~CL\xspace upper limits on the
sparticle production cross section in the SPS8 model, and the NLO cross section
prediction, as a function of $\Lambda$ and the lightest neutralino
and chargino masses. Further SPS8 model parameters
are $M_\mathrm{mess}=2\Lambda$, $\ensuremath{N_{5}}\xspace=1$, $\tan\beta=15$, and
$c\tau_{\mathrm{NLSP}} < \unit[0.1]{mm}$.
\label{fig:sps8:limit}
}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=0.44\textwidth]{result_cls_2011.eps}
\end{center}
\caption{Expected and observed \unit[95]{\%}~CL\xspace upper limits on the
KK particle production cross section times branching fraction to
two photons in the UED model, and the LO
cross section prediction times branching fraction,
as a function of $1/R$ and the KK quark ($Q^{*}$) and KK gluon ($g^{*}$) masses. The UED model parameters are $N=6$, $M_D
= \unit[5]{TeV}$, and $\Lambda R = 20$. }
\label{fig:ued:limit}
\end{figure}
\section{Results}
Based on the observation of 5\xspace events with
$\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace>\unit[125]{GeV}$ and a background expectation of \ensuremath{4.1 \pm 0.6 (\mathrm{stat}) \pm 1.6 (\mathrm{syst})}\xspace
events, a \unit[95]{\%}~CL\xspace upper limit is set on the number of
events in the signal region from any scenario of physics beyond the SM
using the profile likelihood and $CL_s$ method~\cite{Read:2002hq}.
The result is \ensuremath{7.1}\xspace events at \unit[95]{\%}~CL\xspace.
Further, \unit[95]{\%}~CL\xspace upper limits on the cross sections of the
considered models are calculated, including all systematic
uncertainties except for theory uncertainties, i.e. PDF and scale. In
the GGM model the upper limit on the cross section is
\ensuremath{\unit[(22-129)]{fb}}\xspace, where the larger value
corresponds to $\mass{\ensuremath{\tilde{g}}\xspace},\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace} =
\unit[(400,50)]{GeV}$. For $\mass{\ensuremath{\tilde{\chi}^{0}_{1}}\xspace}\ge\unit[150]{GeV}$, the
limit is below \unit[30]{fb}, reaching \unit[22]{fb} for heavy
neutralino masses. Figure~\ref{fig:ggm:limit} shows the expected and
observed lower limits on the GGM gluino mass as a function of the
neutralino mass. For comparison the lower limits from
ATLAS~\cite{Aad:2011kz} and CMS~\cite{Chatrchyan:2011wc} based on the
2010 data are also shown. The total systematic uncertainty includes
the theory uncertainties, which are dominant. Excluding the
PDF and scale uncertainty in the limit calculation would improve the
observed limit on the gluino mass by \unit[$\sim$10]{GeV}.
In the SPS8 model the cross section limit is \ensuremath{\sigma < \unit[(27-91)]{fb}}\xspace as shown in
Fig.\xspace~\ref{fig:sps8:limit}, corresponding to
$\Lambda=\unit[220-80]{TeV}$. For illustration the cross section
dependence as a function of the lightest neutralino and chargino
masses is also shown. A lower limit on the SPS8 breaking scale
$\Lambda > \unit[145]{TeV}\xspace$ at \unit[95]{\%}~CL\xspace is set including the theory
uncertainties, i.e. PDF and scale uncertainties, in the total
systematic uncertainty.
For the UED model the cross section limit is \ensuremath{\sigma < \unit[(15-27)]{fb}}\xspace for
$1/R=1000-\unit[1500]{GeV}$. Figure~\ref{fig:ued:limit} shows the
limit on the cross section times branching ratio for the UED model,
which is \ensuremath{\sigma < \unit[(13-15)]{fb}}\xspace. For illustration the cross section dependence as
a function of the KK quark and KK gluon masses is also shown. A lower
limit on the UED compactification scale $1/R > \unit[1.23]{TeV}\xspace$ at
\unit[95]{\%}~CL\xspace is set.
In this case PDF and scale uncertainties are not included when
calculating the limits. Including PDF and scale uncertainties computed
at LO degrade the limit on $1/R$ by a few \unit{GeV}.
\section{Conclusions}
A search for events with two photons and $\ensuremath{E_{\mathrm{T}}^{\mathrm{miss}}}\xspace>\unit[125]{GeV}$,
performed using \unit[1.07]{\ifb}\xspace of \unit[7]{TeV} $pp$ collision data
recorded with the ATLAS detector at the LHC, found 5\xspace events
with an expected background of \ensuremath{4.1 \pm 0.6 (\mathrm{stat}) \pm 1.6 (\mathrm{syst})}\xspace.
The results are used to set a model-independent \unit[95]{\%}~CL\xspace
upper limit of \ensuremath{7.1}\xspace events from new physics. Upper limits at
\unit[95]{\%}~CL\xspace are also set on the production cross section for
three particular models of new physics: \ensuremath{\sigma < \unit[(22-129)]{fb}}\xspace for the GGM model,
\ensuremath{\sigma < \unit[(27-91)]{fb}}\xspace for the SPS8 model, and \ensuremath{\sigma < \unit[(15-27)]{fb}}\xspace for the UED model.
Under the GGM hypothesis, a lower limit on the gluino mass of
\unit[805]{GeV}\xspace is determined for bino masses above \unit[50]{GeV}. A lower
limit of \unit[145]{TeV}\xspace is set on the SPS8 breaking scale $\Lambda$, which
is the first limit on the SPS8 model at the LHC. A lower limit of
\unit[1.23]{TeV}\xspace is set on the UED compactification scale $1/R$. These
results provide the most stringent tests of these models to date,
significantly improving upon previous best limits of
\unit[560]{GeV}~\cite{Aad:2011kz} for the GGM gluino mass,
\unit[124]{TeV}~\cite{Abazov:2010us} for $\Lambda$ in SPS8, and
\unit[961]{GeV}~\cite{Aad:2011kz} for $1/R$ in UED, respectively.
\section*{Acknowledgements}
\input{Acknowledgement-09Mar11-collisions.tex}
\bibliographystyle{atlasnote}
|
\section{Introduction}
Lithium (Li) is an important diagnostic tool in stellar evolution because its
abundance strongly depends on the ambient conditions. It is quickly destroyed
at $T > 3 \times 10^6$\,K so that it diminishes if the stellar surface is
brought into contact with hot layers by mixing processes. However, if the
overturn time scale for mixing becomes faster than the decay of the parent
$^7$Be, then Li destruction reverts into production through the so-called
Cameron-Fowler mechanism \citep{CF71}. In low-mass stars during the
main-sequence phase, the initial Li abundance strongly decreases, hence slow
mixing should prevail \citep{Mic86}. During the ascent on the red giant branch
(RGB), any Li remaining in the envelope is further diluted by the first
dredge-up (FDU); stellar models excluding atomic diffusion and rotation predict
a surface Li abundance at FDU of $\log\epsilon({\rm Li}) \le +1.5$ (where
$\log\epsilon({\rm Li}) = \log[N({\rm Li})/N({\rm H})]+12$) at this stage. This
is indeed what is observed in most G-K giants \citep{Lam80,Bro89,Mis06}, where
some of these stars show Li even far below the expectations \citep{Mal99}.
However, about 1 -- 2\,\% of the K giants have a high Li abundance and are
therefore called {\em Li-rich} \citep{Bro89,dlR97}.
Advanced evolutionary stages in which Li abundances might possibly increase
through quite fast mixing have been identified by
\citet[][hereafter CB00]{CB00}, in the moments at which the discontinuity of
molecular weight $\mu$ left behind by the downward envelope expansion is erased
by the advancement of the H-burning shell. This occurs at the bump of the
luminosity function, which is on the RGB for low masses. For intermediate mass
stars, this moment is postponed until after core He-burning, when the star
climbs for the second time towards the Hayashi track and the envelope deepens
again (on the so-called asymptotic giant branch, or AGB). The absence of a
$\mu$-barrier is the key factor because it allows any diffusive or plume-like
mixing mechanism to link the envelope with inner layers, and in some cases does
it fast enough, to activate a Cameron-Fowler mechanism for Li production.
However, the proposed location of the enrichment phase directly at the RGB bump
has been questioned by recent observational results \citep[e.g.][]{Mon11}.
Observations of Li in metal-poor globular cluster stars show that a slow
extra-mixing episode at the RGB bump efficiently destroys any Li remaining
after FDU \citep{Lind09b}. Hence, evolved Li-rich stars brighter than the RGB
bump are evidence that those stars somehow skip this mixing episode or even
undergo a phase of Li production, which motivates the search for these stars. As
it is unlikely that they skipped Li destruction at FDU, they must have somehow
replenished it through Be decay \citep[see the discussion in][]{Palm11}.
While the Li abundance has been measured already in an increasing number of
field stars \citep[e.g.][]{Mal99,Kumar11,Mon11} the samples investigated always
had the disadvantage of either inhomogeneity, especially in mass and age,
uncertainty of the distance, or strong selection biases (e.g.\ only stars
around the RGB bump). This circumstance hampers drawing definite conclusions on
the evolutionary phase and length of the Li enrichment. Homogeneous samples,
such as RGB stars in a cluster, are small \citep[e.g.][]{Pas01,Pas04}. We
therefore designed a programme to spectroscopically observe a large and
homogeneous sample of stars from the bottom to the tip of the RGB belonging
to the Galactic bulge (GB) to derive their Li abundances. The GB offers a large
number of low-mass giants at roughly equal distance within a small area in the
sky, perfectly suited for observations with a multi-object spectrograph such as
FLAMES-GIRAFFE. Furthermore, in contrast to globular clusters, the Bulge
contains stars in a wide range of metallicities, which provides the opportunity
to check for any dependence of the Li enrichment processes on this parameter.
We present the results of this programme concerning our search for Li. It will
not only help to confine the phases and duration of Li production on the RGB,
but will also provide a reference value for the Li abundance in the outer GB,
which is needed for interpreting the first Li-rich AGB stars detected there
recently \citep[][hereafter U07]{Utt07}. The present data set is also of high
interest for the study of the structure and nature of the bulge itself, the
results of which will be presented in a forthcoming paper (Uttenthaler et al.,
in preparation).
\section{Sample and observations}\label{obs}
\subsection{Target selection}
The selection of targets was based on data from the 2MASS catalogue
\citep{2MASS} in a 25\arcmin\ diameter circle towards the direction
$(l,b)=(0\degr,-10\degr)$, which is the centre of the Palomar-Groningen field
no.~3 (PG3). This field in the outer GB was chosen because the sample of AGB
stars studied by U07 is also located in the PG3. The GB stars in this field are
roughly 1.4\,kpc away from the Galactic plane and centre. A colour-magnitude
diagram (CMD) to illustrate the target selection is displayed in
Fig.~\ref{cmd}. The observed targets (black circles) were chosen to fall close
to two isochrones from \citet{Gir00}, which are representative for the GB:
$Z=0.004$ and age $10\times10^9$ years, and $Z=0.019$
\citep[which is $Z_{\odot}$ on the scale used by][]{Gir00} and age $5\times10^9$
years. A distance modulus of $14\fm5$ to the GB was adopted. The chosen targets
were allowed to have a $(J - K)_0$ colour either bracketed by the isochrones,
or $0\fm02$ redder or bluer than these. We dereddened the stars using the
linear relation for the reddening in the $B_J$ photographic band as a function
of galactic latitude given in \citet{Schulthe98}. To translate this into the
extinction in the $J$- and $K$-bands, we used the relation $R = A_{V}/E(B- V)$
with $R = 3.2$ and the reddening law of \citet{GS2003}. We find a mean
reddening of $0\fm129$ in the $J$-band and $0\fm049$ in the $K$-band. A
slightly higher reddening is found using the map of \citet{Schleg98}, but the
difference is within the errors of the 2MASS photometry. In addition to the
selection with the colour criterion we selected only stars fainter than
$J_0=9\fm0$ to exclude AGB stars above the tip of the RGB, and stars brighter
than $J_0=14\fm5$ to include the RGB bump, which is expected from isochrones at
$13\fm8 \leqslant J_0 \leqslant 14\fm1$. Furthermore, we discarded all targets
that had fewer than two quality flags 'A' in the 2MASS $JHK$ photometry and
targets that had another source within 3\arcsec\ that was not fainter by at
least $2\fm0$ in the $J$-band than the target itself. This selection scheme
provides a bias against close visual pairs in our sample. Applying all
those criteria yielded 514 targets for the observations.
Also discernible in Fig.~\ref{cmd} are the two red clumps (RCs) recently
identified by \citet{Nat10} and \citet{MZ10}, which probably represent two
populations at different distance in the GB, which are of similar age and
metallicity \citep{DeP11}. These structures were not known to us in the design
phase of the programme. Initially, the fainter RC at $J_0 \sim 13\fm9$ was
interpreted by us to be the RGB bump, and only the brighter one at
$J_0 \sim 13\fm2$ as the red clump. This recent discovery introduces an
uncertainty of a bulge star's location on the RGB of about 0\fm7. The locations
of the RGB bumps of the two populations are then expected to be slightly above
and below the theoretical line marked in Fig.~\ref{cmd}. Given this range of
distance moduli ($\sim14\fm1$ to $\sim14\fm8$) and the photometric errors from
the 2MASS catalogue, we can estimate the expected range of metallicities in our
sample with the help of the isochrones. We find that at the lower brightness
end, stars down to a metallicity of $[{\rm M}/{\rm H}]=-1.3$ would still fall
in the selection range, albeit with a lower probability than stars with a
metallicity between that of the two selection isochrones. At brighter
magnitudes, the lower metallicity cut will be somewhat higher, but stars down to
$[{\rm M}/{\rm H}]=-0.7$ should be complete in the whole magnitude range of our
sample. On the other hand, stars belonging to the near side of the X-shaped
bulge may still fall within our selection region up to a metallicity of
$[{\rm M}/{\rm H}]=+0.2$. As we will see in Sect.~\ref{analysis}, a few stars
fall outside this range, but those cases should be rare.
\begin{figure}
\centering
\includegraphics[width=\linewidth,bb=87 366 538 700,clip]{CMD.pdf}
\caption{2MASS Colour-magnitude diagram of the field around
$(l,b)=(0\degr,-10\degr)$. Stars not observed in the present programme are
plotted as black dots, observed targets as black circles, stars moderately
enriched in Li as small red circles, and Li-rich stars as big red circles.
Two RGB isochrones from \citet{Gir00}, with metallicities and ages as
indicated in the legend, are also included. The isochrones are truncated at
the tip of the RGB.}
\label{cmd}
\end{figure}
\subsection{Observations and data reduction}
To measure the Li abundance we obtained spectra of the sample stars using the
FLAMES spectrograph in the GIRAFFE configuration. The grating HR15 centred at
665\,nm was used, which gave a spectral resolution of 17\,000 for a wavelength
coverage from 644 to 682\,nm. This setting contains the Li resonance line at
670.8\,nm and the H$\alpha$ line, in addition to several other atomic and
molecular features. Because our sample stretches over a brightness range of 5
magnitudes, it was divided into three sub-groups, a bright
(\mbox{$11\fm9 \lesssim R \lesssim 13\fm3$}), an intermediate
(\mbox{$13\fm3 \lesssim R \lesssim 14\fm7$}), and a faint group
(\mbox{$14\fm7 \lesssim R \lesssim 16\fm2$}), with total exposure times
of 0.5, 1, and 4\,h, respectively\footnote{The $R$-band limits were determined
from the isochrone predictions and used in the GIRAFFE exposure time
calculator.}. For the bright group one fibre configuration was sufficient, while
the larger number of targets at fainter magnitudes made necessary to split up
the intermediate and faint group into two and three fibre configuration,
respectively.
The observations were obtained in service mode in June 2009. To keep observing
blocks below one hour and to handle cosmic ray hits, the observations of the
intermediate group were split into two, those of the faint group into five
exposures. Individual spectra were extracted from each of these images
separately using the FLAMES pipeline provided by ESO. The quality of each
spectrum was checked, and those with low signal or cosmic ray hits in the
critical wavelength ranges were rejected. Before averaging the good spectra,
small wavelength shifts caused by splitting the individual observations in some
cases over a few weeks were corrected. The applied shift agreed with the
expected value for the difference in heliocentric correction, and the same
shift was found for all spectra obtained at a given time. Therefore, we can
exclude that these shifts are the result of orbital motion in a binary system.
No other velocity shifts were found within the accuracy limits of our data.
However, for several stars only one observation was available, therefore we
point out that with the present data set an effective discrimination against
binaries is not possible. We aimed at a signal-to-noise ratio (S/N) close to
100, which was achieved for most targets. Of the initially proposed sample, 401
spectra had sufficiently high quality to check for the presence of the Li line
and to measure their radial velocity (RV).
\subsection{Foreground contamination}
An important point for this kind of study is the foreground contamination of
the sample. To reliably estimate the fraction of genuine bulge stars in our
sample, we ran two simulations of Galactic population models, namely the
Besan\c{c}on model \citep{Rob03} and the TRILEGAL model \citep{Gir05}. For the
TRILEGAL simulation (kindly provided by E.\ Vanhollebeke), the best-fit
parameters of a tri-axial bulge found by \citet{Van09} were adopted. We applied
the same colour and magnitude selection criteria to the simulated data as to
the observed data. This selection yielded 81\% and 74\% bulge stars in the
TRILEGAL and Besan\c{c}on model, respectively. We also inspected the
distribution of initial masses of the simulated stars in the selection area.
The mass spectrum in the TRILEGAL simulation is relatively sharply peaked at
$1.01 M_{\sun}$, with a standard deviation of $0.13 M_{\sun}$. Hence, we are
confident that our sample indeed contains mostly low-mass ($M \sim 1 M_{\sun}$)
red giant stars that are located in the GB.
Some of the foreground stars might be identified by their high proper motion.
We therefore searched the Southern Proper Motion Program catalogue version 4
\citep[SPM4;][]{Gir11} for sample stars with high proper motion. All but ten of
our sample stars were identified in the SPM4 catalogue. Unfortunately, the
uncertainties in the SPM4 catalogue are large, so that bulge and foreground
stars can hardly be separated. We decided to assign those stars to the
foreground whose proper motion is higher than 20\,mas/yr. Applying this
criterion, we found 50 foreground star candidates; these stars are marked with
an asterisk in Table~\ref{onltab}. One star in our sample (\#300) has a
particularly high proper motion of more than 200\,mas/yr, hence it is probably
a nearby K dwarf.
\section{Analysis and results}\label{analysis}
For the following analysis we used hydrostatic COMARCS atmospheres in
connection with the COMA spectral synthesis package \citep{Ari09}. This ensures
a consistent treatment of radiative transfer and opacities during the complete
analysis. We started with a temperature calibration based on 2MASS $(J-K_S)_0$
colours, which is presented in Appendix~\ref{JKTeffsect}. A determination of
the stellar temperature and surface gravity using only the FLAMES spectra was
not possible because of the limited spectral range and the dominance of TiO
lines in the cooler targets. After fixing the temperature based on the colours,
we obtained the surface gravities $(g)$ from isochrones in Fig.~\ref{cmd}
\citep{Gir00}, assuming that the stars belong to the Bulge RGB. The uncertainty
in the distance modulus of $\pm0\fm35$ (Sect.~\ref{obs}) introduces an
uncertainty in $\log(g)$ of $\pm0.16$, which has only a very minor impact on
the uncertainty in the Li abundance, however.
In a first step of the analysis of the spectra, we determined the RVs of the
stars by cross-correlating the observed spectra with a synthetic spectrum that
was based on a COMARCS model with a temperature close to that estimated for
each star. In the next step, the observed spectra were inspected visually
around the 670.8\,nm Li line, together with a synthetic spectrum calculated
with no Li to check for excess absorption. The Li line was found to be
detectable in 31 stars. Four of these Li-bearing stars are also foreground star
candidates (Sect.~\ref{obs}).
To estimate the metallicity of each star, we calculated model atmospheres with
five different metallicities ($[{\rm M}/{\rm H}]=-1.5$, $-1.0$, $-0.5$, $0.0$,
and $+0.5$), with $T_{{\rm eff}}$ and $\log g$ values determined in the previous
step. A general $\alpha$-element enhancement of $[\alpha/{\rm Fe}]=+0.2$, a C/O
ratio of 0.3, and a micro-turbulent velocity typical for red giant stars of
$\xi=2.5$\,km\,s$^{-1}$ were assumed. The spectra with the five different
metallicities were interpolated and fitted to the observed spectra in the
wavelength range 649 -- 680\,nm using the downhill simplex IDL routine
amoeba.pro. The aim of this exercise was not to determine a very precise
metallicity of the stars, but instead to reasonably model the quasi-continuum in
the vicinity of the Li line, which is suppressed from the true continuum by a
forest of weak lines, and the lines blending with the Li line. This procedure
resulted in very satisfying fits of the spectra. The metallicity determined in
this way is probably quite reliable, its accuracy is mainly limited by the
accuracy of the temperature determination. A check of this method on the
observed spectrum of Arcturus yields a metallicity of
$[{\rm M}/{\rm H}]=-0.69$. However, if model atmospheres without an
alpha-enhancement are used, this would be increased to $-0.64$, which
reasonably agrees with the range of metallicities found for this star in the
Simbad database. We find a mean metallicity of $[{\rm M}/{\rm H}]=-0.48$ for
the 31 Li-bearing stars, which is somewhat lower than what would be expected
from previous investigations of Bulge fields close to ours
\citep{Zoc08}\footnote{Performing the averaging on a {\em linear} scale, which
is the mathematically correct way, the mean is $[{\rm M}/{\rm H}]=-0.21$.}.
Some of the stars have very weak metal lines, indeed.
Possibly, there are selection effects introduced in the sense that the Li-rich
phenomenon could be more common among metal-poor stars, and/or that the
detection threshold of Li is lower at lower metallicity because of the reduced
background of blending lines. Note that the three most Li-rich stars (see below)
have a metallicity between $-0.83$ and $-0.96$\,dex below solar.
With the stellar parameters fixed in this way, we synthesised spectra, assuming
varying Li abundances to fit the observed spectra. Atomic line data for the
spectral synthesis were taken from the VALD data base \citep{Kup99}. The line
list was checked by comparing a model spectrum of Arcturus
\citep[adopting the stellar parameters and abundances found by][]{Ryd10} with
the observed Arcturus spectrum \citep{Hin00} in the vicinity of the 671\,nm Li
line. Some lines present in VALD were found to be too strong in the model
spectrum; their $\log(gf)$ were decreased accordingly, or the line was
completely removed from the list. The abundances given in \citet{Caf09} were
adopted as solar reference. We did not apply NLTE corrections to the Li
abundances \citep{Lind09a} derived under LTE assumption because these
corrections are usually small and, for the parameters covered by our sample
stars, within the uncertainties.
The programme stars \#042, \#080, and \#123 are identified to be Li-rich
according to the classical definition, one of them (\#042) has a Li abundance
that agrees with the cosmic value and should be regarded as {\em super}
Li-rich. The observed spectrum of star \#042 around the Li line together with
a synthetic spectrum with the best fitting abundances is displayed in
Fig.~\ref{tatltuae}. The spectrum of a Li-detected star (\#030,
$\log\epsilon({\rm Li})=+0.1$) and of a Li-poor star (\#051) are also shown in
that figure. All other stars with a detected Li line have an abundance at or
below the solar abundance. The error in the Li abundance was estimated by
varying the stellar parameters by their uncertainty. The temperature has the
highest influence on the strength of the Li line, and is uncertain by
$\pm160$\,K for the hotter stars, and $\pm90$\,K for the cooler stars, based on
the uncertainty in the 2MASS photometry. We consequently estimate an
uncertainty in the Li abundance of $\pm0.3$\,dex. Moreover, the detection
threshold varies with the temperature because Li becomes more and more ionised
at higher temperature. A detection threshold of $\log\epsilon({\rm Li})=+0.5$
was derived for the hottest sample stars (4800\,K), which declines to 0.0 at
around 4200\,K, and to $\sim-0.5$ at 3600\,K\footnote{At still lower
temperatures, the detection threshold rises again because of the increasing
strength of the TiO lines.}.
All measured Li abundances as well as important quantities of the targets are
collected in Table~\ref{onltab}. Our results are also illustrated in
Fig.~\ref{cmd}: Stars with a detected Li line are plotted as small red circles,
and the Li-rich stars are plotted as large red circles.
\begin{figure}
\centering
\includegraphics[width=\linewidth,bb=82 369 538 699,clip]{tatltuae.pdf}
\caption{Observed spectra of the the programme stars \#042, the most Li-rich
star in our sample; \#030, a Li-detected star; and \#051, a Li-poor star
(black dots, from bottom to top). For clarity, the spectra of stars \#030
and \#051 have been shifted by +0.2 and +0.4 in flux. The continuous line
is the best-fit synthetic spectrum to star \#042 with a Li abundance of
$\log \epsilon({\rm Li})=+3.2$. The two dotted vertical lines indicate the
laboratory wavelengths of the hyperfine transitions in $^7$Li. Note also
the good fit to two Fe lines at \mbox{$\sim670.55$} and $\sim671.21$\,nm,
which indicates a well-determined me\-tal\-li\-ci\-ty of
$[{\rm M}/{\rm H}]=-0.85$ for star \#042. The metallicities for the two
other stars were determined to be $-0.94$ (\#030) and $-1.03$ (\#051),
respectively.
}
\label{tatltuae}
\end{figure}
\onllongtab{1}{
\begin{longtable}{lcccccrccccc}
\caption{Parameters of the stars and measured abundances.}\label{onltab}\\
\hline
\hline
No.\ & RA(J2000) & Dec(J2000) & $J$ & $H$ & $K$ & RV & $T_{\rm eff}$ & $\log g$ & [M/H] & $\log\epsilon({\rm Li})$ & FG\\
& deg.\ & deg.\ & & & & km\,s$^{-1}$ & K & & & \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) & (10) & (11) & (12) \\
\hline
\endfirsthead
\caption{Continued.}\\
\hline
\hline
No.\ & RA(J2000) & Dec(J2000) & $J$ & $H$ & $K$ & RV & $T_{\rm eff}$ & $\log g$ & [M/H] & $\log\epsilon({\rm Li})$ & FG\\
& deg.\ & deg.\ & & & & km\,s$^{-1}$ & K & & & \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) & (10) & (11) & (12) \\
\hline
\endhead
\hline
\endfoot
\hline
\endlastfoot
001 & 276.814513 & $-33.589966$ & 9.495 & 8.524 & 8.256 & 117.328 & 3465 & & & & *\\
002 & 276.544710 & $-33.787083$ & 9.512 & 8.624 & 8.369 & 20.277 & 3697 & & & & \\
003 & 276.608107 & $-33.667065$ & 9.516 & 8.615 & 8.298 & 76.484 & 3520 & & & & \\
005 & 276.918684 & $-33.751503$ & 9.696 & 8.820 & 8.527 & 55.090 & 3629 & & & & \\
006 & 276.432104 & $-33.784958$ & 9.926 & 9.061 & 8.704 & $ -29.587$ & 3514 & & & & \\
007 & 276.714893 & $-33.891556$ & 9.978 & 9.096 & 8.838 & 12.913 & 3698 & & & & \\
008 & 276.600416 & $-33.681347$ & 10.007 & 9.121 & 8.891 & $-135.108$ & 3756 & 0.5 &$-0.14$& 0.4 & \\
009 & 276.807769 & $-33.757442$ & 10.057 & 9.178 & 8.941 & 106.494 & 3749 & & & & \\
010 & 276.873765 & $-33.726830$ & 10.098 & 9.183 & 8.932 & $-128.768$ & 3637 & 0.6 &$-0.59$&$-0.3$ & \\
011 & 276.817925 & $-33.649303$ & 10.115 & 9.191 & 8.923 & $ -59.147$ & 3578 & & & & \\
012 & 276.611791 & $-33.631435$ & 10.170 & 9.330 & 9.094 & 74.703 & 3843 & & & & \\
013 & 276.506457 & $-33.723846$ & 10.213 & 9.322 & 9.062 & 27.664 & 3681 & & & & \\
015 & 276.488172 & $-33.844830$ & 10.232 & 9.400 & 9.158 & 85.083 & 3847 & & & & \\
016 & 276.682337 & $-33.955391$ & 10.266 & 9.421 & 9.194 & 25.704 & 3845 & & & & \\
017 & 276.595604 & $-33.694199$ & 10.406 & 9.574 & 9.312 & 50.134 & 3803 & & & & \\
018 & 276.592533 & $-33.625336$ & 10.409 & 9.579 & 9.366 & $ -59.013$ & 3914 & & & & \\
019 & 276.586978 & $-33.867764$ & 10.470 & 9.639 & 9.438 & 178.014 & 3934 & & & & \\
020 & 276.704314 & $-33.794868$ & 10.542 & 9.678 & 9.446 & 33.422 & 3794 & & & & *\\
021 & 276.653399 & $-33.955818$ & 10.642 & 9.747 & 9.556 & 72.019 & 3815 & & & & \\
022 & 276.471245 & $-33.828197$ & 10.657 & 9.769 & 9.520 & $ -80.548$ & 3711 & 0.9 &$+0.08$& 0.5 & *\\
023 & 276.565931 & $-33.660713$ & 10.690 & 9.821 & 9.542 & $ -42.688$ & 3687 & & & & \\
024 & 276.667873 & $-33.869152$ & 10.726 & 9.949 & 9.659 & $ -7.078$ & 3857 & & & & \\
025 & 276.771326 & $-33.748791$ & 10.735 & 9.924 & 9.721 & $-131.548$ & 3969 & & & & \\
026 & 276.566599 & $-33.576355$ & 10.747 & 9.916 & 9.673 & 15.044 & 3849 & & & & \\
027 & 276.621730 & $-33.925117$ & 10.779 & 9.943 & 9.708 & $ -0.464$ & 3849 & 0.9 &$+0.11$& 0.6 & \\
028 & 276.566492 & $-33.894100$ & 10.807 & 10.005 & 9.774 & $ -78.865$ & 3932 & & & & \\
029 & 276.632910 & $-33.850189$ & 10.843 & 10.032 & 9.847 & 305.574 & 4008 & & & & \\
030 & 276.876718 & $-33.749680$ & 10.911 & 10.121 & 9.909 & 13.784 & 3991 & 1.0 &$-0.94$& 0.1 & \\
031 & 276.833702 & $-33.646965$ & 10.933 & 10.095 & 9.844 & $ -12.983$ & 3808 & & & & \\
032 & 276.664170 & $-33.889168$ & 11.002 & 10.132 & 9.895 & 59.999 & 3770 & & & & \\
033 & 276.608510 & $-33.775723$ & 11.064 & 10.240 & 10.041 & 46.289 & 3953 & & & & \\
034 & 276.530697 & $-33.808578$ & 11.077 & 10.243 & 10.056 & 72.605 & 3959 & & & & \\
035 & 276.734398 & $-33.905167$ & 11.112 & 10.288 & 10.008 & 37.840 & 3775 & & & & \\
036 & 276.693013 & $-33.893147$ & 11.119 & 10.247 & 10.060 & 110.060 & 3874 & & & & \\
037 & 276.455229 & $-33.789265$ & 11.126 & 10.329 & 10.132 & 25.919 & 4017 & & & & \\
038 & 276.531016 & $-33.587563$ & 11.157 & 10.323 & 10.122 & $-105.342$ & 3933 & & & & \\
039 & 276.476724 & $-33.652061$ & 11.191 & 10.461 & 10.223 & $ -99.668$ & 4074 & 1.2 &$-0.17$& 0.3 & \\
040 & 276.632713 & $-33.608341$ & 11.204 & 10.397 & 10.238 & $-213.279$ & 4075 & & & & \\
041 & 276.611649 & $-33.623035$ & 11.212 & 10.433 & 10.228 & 7.052 & 4037 & & & & \\
042 & 276.514543 & $-33.867790$ & 11.212 & 10.396 & 10.256 & 3.906 & 4096 & 1.2 &$-0.85$& 3.2 & \\
043 & 276.545896 & $-33.910049$ & 11.236 & 10.429 & 10.287 & $ -60.971$ & 4109 & 1.2 &$-0.61$& 0.3 & \\
044 & 276.501519 & $-33.712631$ & 11.270 & 10.456 & 10.228 & $ -66.401$ & 3917 & & & & \\
045 & 276.497877 & $-33.881786$ & 11.378 & 10.535 & 10.362 & 240.911 & 3969 & & & & \\
046 & 276.469519 & $-33.651062$ & 11.403 & 10.619 & 10.362 & $ -1.719$ & 3921 & & & & \\
047 & 276.718276 & $-33.890141$ & 11.407 & 10.593 & 10.388 & 29.551 & 3957 & & & & \\
048 & 276.593535 & $-33.598366$ & 11.415 & 10.584 & 10.345 & $ -9.884$ & 3857 & & & & \\
049 & 276.751404 & $-33.686954$ & 11.428 & 10.661 & 10.459 & 81.687 & 4065 & & & & \\
050 & 276.668148 & $-33.844818$ & 11.433 & 10.578 & 10.353 & 29.158 & 3829 & & & & \\
051 & 276.897946 & $-33.802887$ & 11.446 & 10.612 & 10.445 & $ -53.729$ & 3992 & & & & \\
052 & 276.827522 & $-33.672203$ & 11.457 & 10.650 & 10.435 & $ -28.859$ & 3952 & 1.4 &$+0.09$& 0.4 & \\
053 & 276.598353 & $-33.763283$ & 11.464 & 10.658 & 10.412 & 16.198 & 3893 & & & & \\
054 & 276.753347 & $-33.748871$ & 11.480 & 10.655 & 10.504 & 21.622 & 4049 & 1.2 &$-0.94$& & \\
058 & 276.763338 & $-33.726757$ & 11.555 & 10.775 & 10.634 & $-133.305$ & 4167 & & & & \\
060 & 276.814847 & $-33.799526$ & 11.592 & 10.814 & 10.653 & 19.204 & 4126 & & & & \\
061 & 276.745468 & $-33.614552$ & 11.610 & 10.791 & 10.615 & $-206.421$ & 4011 & & & & \\
065 & 276.767964 & $-33.723507$ & 11.682 & 10.895 & 10.664 & $ -59.417$ & 3961 & & & & \\
067 & 276.764757 & $-33.714439$ & 11.695 & 10.914 & 10.713 & $ -42.029$ & 4036 & & & & \\
069 & 276.732436 & $-33.691238$ & 11.750 & 10.953 & 10.767 & 28.948 & 4035 & & & & \\
070 & 276.809016 & $-33.594540$ & 11.763 & 10.955 & 10.743 & $ -54.899$ & 3957 & 1.4 &$-0.39$& 0.2 & \\
072 & 276.594486 & $-33.730743$ & 11.791 & 11.111 & 10.937 & $ -52.206$ & 4325 & & & & \\
073 & 276.836739 & $-33.717419$ & 11.816 & 11.071 & 10.883 & $ -56.535$ & 4139 & & & & \\
075 & 276.910260 & $-33.767544$ & 11.837 & 11.034 & 10.873 & 45.375 & 4070 & & & & *\\
080 & 276.663676 & $-33.680828$ & 11.956 & 11.185 & 11.002 & 67.611 & 4099 & 1.5 &$-0.83$& 2.0 & \\
081 & 276.598738 & $-33.673889$ & 11.964 & 11.282 & 11.096 & $-155.581$ & 4294 & 1.5 &$-0.32$& 0.7 & \\
084 & 276.723840 & $-33.562702$ & 11.989 & 11.339 & 11.120 & 29.832 & 4290 & & & & \\
085 & 276.663499 & $-33.799038$ & 12.006 & 11.199 & 11.069 & 58.329 & 4134 & & & & *\\
086 & 276.810910 & $-33.617214$ & 12.009 & 11.269 & 11.096 & 9.035 & 4185 & & & & \\
088 & 276.843442 & $-33.804676$ & 12.058 & 11.340 & 11.215 & 84.524 & 4343 & & & & \\
089 & 276.708411 & $-33.645660$ & 12.060 & 11.331 & 11.179 & $ -58.044$ & 4261 & & & & \\
096 & 276.728929 & $-33.625156$ & 12.172 & 11.457 & 11.252 & 38.482 & 4171 & & & & \\
101 & 276.650041 & $-33.653530$ & 12.257 & 11.553 & 11.355 & $-123.588$ & 4213 & & & & \\
102 & 276.603598 & $-33.725780$ & 12.263 & 11.594 & 11.421 & $ -37.926$ & 4352 & & & & \\
104 & 276.764294 & $-33.611069$ & 12.270 & 11.528 & 11.326 & 23.454 & 4119 & & & & \\
107 & 276.796933 & $-33.752850$ & 12.304 & 11.596 & 11.481 & 16.073 & 4392 & & & & \\
108 & 276.818554 & $-33.721001$ & 12.315 & 11.546 & 11.354 & $ -55.377$ & 4080 & & & & \\
110 & 276.831116 & $-33.661152$ & 12.343 & 11.625 & 11.495 & $-109.396$ & 4334 & & & & \\
112 & 276.844085 & $-33.719276$ & 12.386 & 11.681 & 11.584 & $ -3.301$ & 4444 & & & & \\
115 & 276.876631 & $-33.669006$ & 12.398 & 11.613 & 11.454 & $-172.477$ & 4115 & 1.7 &$-0.56$&$<-0.3$& \\
116 & 276.886129 & $-33.817890$ & 12.401 & 11.716 & 11.560 & $ -54.145$ & 4346 & & & & *\\
117 & 276.839217 & $-33.735954$ & 12.407 & 11.626 & 11.535 & $-141.307$ & 4277 & & & & \\
118 & 276.605428 & $-33.607613$ & 12.438 & 11.750 & 11.607 & 31.103 & 4380 & 1.7 &$-1.19$&$<-0.3$& \\
121 & 276.850407 & $-33.727764$ & 12.470 & 11.850 & 11.667 & $ -2.194$ & 4441 & & & & \\
123 & 276.733228 & $-33.686802$ & 12.482 & 11.722 & 11.609 & $ -85.450$ & 4278 & 1.7 &$-0.96$& 2.1 & \\
128 & 276.674814 & $-33.794342$ & 12.520 & 11.823 & 11.717 & 32.514 & 4446 & & & & \\
131 & 276.862902 & $-33.842964$ & 12.596 & 11.871 & 11.725 & $ -28.810$ & 4277 & & & & \\
132 & 276.529376 & $-33.635666$ & 12.607 & 11.953 & 11.827 & $-159.866$ & 4514 & & & & \\
133 & 276.754996 & $-33.640835$ & 12.630 & 11.933 & 11.784 & 136.734 & 4341 & & & & *\\
136 & 276.670396 & $-33.678337$ & 12.694 & 11.991 & 11.848 & $ -16.510$ & 4342 & & & & \\
139 & 276.669257 & $-33.560745$ & 12.724 & 12.024 & 11.849 & $ -8.143$ & 4278 & & & & \\
142 & 276.592785 & $-33.609268$ & 12.796 & 12.171 & 11.994 & $-120.627$ & 4454 & & & & \\
143 & 276.720586 & $-33.791100$ & 12.801 & 12.219 & 11.977 & $ -28.694$ & 4391 & & & & *\\
144 & 276.724865 & $-33.593685$ & 12.806 & 12.156 & 12.024 & 47.662 & 4504 & & & & \\
147 & 276.605327 & $-33.599644$ & 12.834 & 12.130 & 12.014 & $ -13.128$ & 4407 & & & & \\
148 & 276.808190 & $-33.749828$ & 12.835 & 12.127 & 11.988 & $ -40.138$ & 4335 & & & & \\
154 & 276.884497 & $-33.711712$ & 12.903 & 12.231 & 12.130 & $ -74.863$ & 4521 & & & & \\
156 & 276.802245 & $-33.615543$ & 12.915 & 12.209 & 12.115 & 5.379 & 4453 & & & & *\\
158 & 276.607949 & $-33.626446$ & 12.936 & 12.308 & 12.179 & 47.731 & 4574 & & & & \\
159 & 276.506642 & $-33.859463$ & 12.947 & 12.320 & 12.168 & $ -67.344$ & 4514 & & & & \\
167 & 276.696311 & $-33.687943$ & 12.992 & 12.354 & 12.170 & 16.036 & 4398 & & & & \\
169 & 276.585971 & $-33.583626$ & 13.016 & 12.356 & 12.206 & $ -21.109$ & 4434 & & & & \\
172 & 276.674085 & $-33.722809$ & 13.036 & 12.399 & 12.232 & 62.310 & 4444 & & & & \\
178 & 276.673132 & $-33.740860$ & 13.069 & 12.451 & 12.242 & $ -14.891$ & 4385 & & & & \\
180 & 276.715472 & $-33.633411$ & 13.078 & 12.449 & 12.266 & 57.926 & 4424 & & & & *\\
182 & 276.772886 & $-33.712326$ & 13.085 & 12.412 & 12.316 & $ -44.163$ & 4535 & & & & \\
183 & 276.791120 & $-33.629635$ & 13.088 & 12.478 & 12.337 & $ -74.273$ & 4584 & & & & \\
184 & 276.841313 & $-33.728474$ & 13.093 & 12.437 & 12.271 & 13.471 & 4393 & & & & \\
186 & 276.721044 & $-33.780018$ & 13.110 & 12.489 & 12.371 & 29.773 & 4618 & & & & \\
187 & 276.533345 & $-33.873089$ & 13.112 & 12.435 & 12.314 & $-142.788$ & 4462 & & & & \\
188 & 276.753476 & $-33.945801$ & 13.119 & 12.543 & 12.370 & 19.007 & 4585 & & & & \\
190 & 276.538801 & $-33.644936$ & 13.134 & 12.484 & 12.360 & 11.192 & 4530 & 2.1 &$+0.27$& 0.5 & \\
191 & 276.627097 & $-33.841202$ & 13.146 & 12.522 & 12.389 & $-175.576$ & 4569 & & & & \\
192 & 276.790756 & $-33.846645$ & 13.151 & 12.493 & 12.363 & $ -8.106$ & 4481 & & & & \\
193 & 276.717695 & $-33.835640$ & 13.153 & 12.548 & 12.410 & 56.213 & 4605 & & & & \\
194 & 276.864580 & $-33.723576$ & 13.154 & 12.547 & 12.417 & $ -33.430$ & 4620 & & & & \\
195 & 276.694217 & $-33.601109$ & 13.163 & 12.508 & 12.393 & $ -3.184$ & 4537 & & & & \\
196 & 276.841761 & $-33.689163$ & 13.167 & 12.539 & 12.385 & $ -44.337$ & 4499 & & & & *\\
197 & 276.749744 & $-33.856068$ & 13.171 & 12.572 & 12.406 & $ -22.102$ & 4544 & & & & \\
198 & 276.638867 & $-33.930489$ & 13.179 & 12.505 & 12.351 & 89.036 & 4381 & & & & \\
199 & 276.601595 & $-33.643906$ & 13.190 & 12.594 & 12.394 & 55.281 & 4469 & & & & \\
200 & 276.910018 & $-33.774948$ & 13.190 & 12.585 & 12.420 & $ -23.265$ & 4527 & & & & \\
201 & 276.655418 & $-33.667728$ & 13.192 & 12.617 & 12.458 & $ -12.106$ & 4637 & & & & \\
202 & 276.600953 & $-33.920620$ & 13.193 & 12.561 & 12.403 & $ -72.007$ & 4480 & & & & \\
203 & 276.718361 & $-33.917198$ & 13.194 & 12.505 & 12.362 & $-112.531$ & 4370 & & & & \\
204 & 276.651481 & $-33.734238$ & 13.196 & 12.496 & 12.354 & $ -44.013$ & 4351 & & & & \\
205 & 276.778831 & $-33.720421$ & 13.203 & 12.613 & 12.452 & $ -66.094$ & 4583 & & & & \\
206 & 276.529719 & $-33.594994$ & 13.210 & 12.613 & 12.466 & 26.825 & 4613 & & & & \\
207 & 276.854131 & $-33.731907$ & 13.221 & 12.595 & 12.447 & 62.408 & 4519 & & & & \\
208 & 276.628729 & $-33.743752$ & 13.226 & 12.621 & 12.494 & $ -39.333$ & 4642 & & & & \\
209 & 276.798530 & $-33.662918$ & 13.233 & 12.540 & 12.415 & 95.619 & 4405 & & & & \\
210 & 276.836967 & $-33.769405$ & 13.243 & 12.643 & 12.456 & 21.124 & 4484 & & & & \\
211 & 276.756279 & $-33.788357$ & 13.248 & 12.575 & 12.430 & $ -34.348$ & 4404 & & & & \\
212 & 276.892143 & $-33.753590$ & 13.248 & 12.670 & 12.487 & 21.790 & 4552 & & & & *\\
215 & 276.537753 & $-33.603622$ & 13.256 & 12.649 & 12.491 & $-119.873$ & 4555 & & & & *\\
216 & 276.630558 & $-33.724087$ & 13.258 & 12.652 & 12.484 & 38.871 & 4526 & & & & \\
217 & 276.611185 & $-33.719536$ & 13.258 & 12.650 & 12.506 & 53.834 & 4586 & & & & \\
218 & 276.533293 & $-33.793808$ & 13.263 & 12.630 & 12.449 & 25.411 & 4421 & & & & \\
219 & 276.490017 & $-33.791950$ & 13.267 & 12.645 & 12.524 & 70.196 & 4614 & & & & \\
220 & 276.786346 & $-33.598869$ & 13.267 & 12.685 & 12.496 & 36.370 & 4531 & & & & *\\
221 & 276.488605 & $-33.676846$ & 13.277 & 12.662 & 12.508 & $ -68.765$ & 4544 & & & & \\
222 & 276.765113 & $-33.566185$ & 13.280 & 12.582 & 12.447 & 89.637 & 4371 & & & & \\
223 & 276.689571 & $-33.850536$ & 13.283 & 12.596 & 12.483 & 63.721 & 4452 & 2.1 &$-0.01$& 0.5 & \\
224 & 276.530160 & $-33.772533$ & 13.283 & 12.630 & 12.546 & $-159.749$ & 4630 & & & & *\\
225 & 276.490013 & $-33.854095$ & 13.287 & 12.654 & 12.528 & $ -2.227$ & 4568 & & & & \\
226 & 276.441487 & $-33.833378$ & 13.287 & 12.601 & 12.440 & 86.733 & 4344 & & & & \\
227 & 276.460830 & $-33.865948$ & 13.295 & 12.693 & 12.545 & $ -30.132$ & 4593 & & & & \\
228 & 276.567606 & $-33.881664$ & 13.305 & 12.729 & 12.573 & 5.293 & 4641 & & & & \\
229 & 276.488238 & $-33.744823$ & 13.308 & 12.659 & 12.507 & $ -93.491$ & 4457 & & & & \\
230 & 276.701957 & $-33.724052$ & 13.317 & 12.670 & 12.492 & $ -15.167$ & 4390 & & & & *\\
231 & 276.521504 & $-33.882675$ & 13.318 & 12.634 & 12.517 & 87.725 & 4454 & & & & \\
232 & 276.701163 & $-33.812206$ & 13.321 & 12.791 & 12.593 & 46.717 & 4650 & & & & \\
233 & 276.742927 & $-33.655430$ & 13.327 & 12.768 & 12.577 & $ -33.817$ & 4588 & & & & \\
234 & 276.817331 & $-33.599476$ & 13.328 & 12.681 & 12.582 & 3.010 & 4598 & & & & \\
235 & 276.561797 & $-33.754608$ & 13.331 & 12.760 & 12.590 & 2.381 & 4618 & & & & \\
236 & 276.811041 & $-33.747356$ & 13.339 & 12.698 & 12.552 & $ -36.405$ & 4485 & & & & *\\
237 & 276.543824 & $-33.638008$ & 13.349 & 12.771 & 12.601 & $-128.142$ & 4601 & & & & \\
238 & 276.707179 & $-33.863255$ & 13.363 & 12.721 & 12.560 & $ -48.485$ & 4443 & & & & \\
239 & 276.514484 & $-33.893902$ & 13.367 & 12.706 & 12.596 & $ -6.349$ & 4534 & & & & \\
240 & 276.494178 & $-33.865170$ & 13.374 & 12.671 & 12.545 & $ -45.410$ & 4384 & & & & \\
241 & 276.691228 & $-33.652210$ & 13.381 & 12.740 & 12.621 & 163.130 & 4563 & & & & \\
242 & 276.769980 & $-33.622349$ & 13.400 & 12.785 & 12.587 & 37.049 & 4420 & & & & \\
243 & 276.445539 & $-33.820004$ & 13.401 & 12.802 & 12.662 & $-201.762$ & 4626 & & & & *\\
245 & 276.458400 & $-33.680145$ & 13.410 & 12.870 & 12.674 & $ -0.726$ & 4637 & & & & \\
246 & 276.732205 & $-33.735016$ & 13.413 & 12.722 & 12.633 & $ -20.965$ & 4506 & & & & \\
247 & 276.652760 & $-33.666142$ & 13.415 & 12.859 & 12.647 & $ -12.734$ & 4542 & & & & \\
248 & 276.663722 & $-33.563961$ & 13.423 & 12.782 & 12.703 & $ -91.605$ & 4680 & & & & \\
249 & 276.874535 & $-33.706005$ & 13.432 & 12.805 & 12.675 & 20.002 & 4564 & & & & \\
251 & 276.814218 & $-33.891129$ & 13.436 & 12.813 & 12.703 & $-100.233$ & 4630 & & & & \\
252 & 276.590433 & $-33.824047$ & 13.447 & 12.815 & 12.625 & 14.280 & 4399 & & & & \\
253 & 276.698884 & $-33.720577$ & 13.450 & 12.802 & 12.707 & $ -12.354$ & 4608 & & & & \\
254 & 276.560443 & $-33.786304$ & 13.457 & 12.905 & 12.659 & 47.473 & 4463 & & & & \\
256 & 276.885368 & $-33.847267$ & 13.459 & 12.843 & 12.729 & 47.665 & 4637 & & & & \\
257 & 276.673226 & $-33.739399$ & 13.464 & 12.896 & 12.744 & 53.227 & 4676 & & & & \\
258 & 276.509126 & $-33.638287$ & 13.466 & 12.795 & 12.648 & 20.249 & 4414 & & & & \\
259 & 276.903426 & $-33.794418$ & 13.475 & 12.812 & 12.641 & 0.338 & 4362 & & & & \\
260 & 276.856050 & $-33.636570$ & 13.477 & 12.791 & 12.647 & $ -99.944$ & 4375 & & & & \\
261 & 276.597652 & $-33.706051$ & 13.478 & 12.889 & 12.705 & $ -56.387$ & 4530 & & & & \\
262 & 276.627210 & $-33.776421$ & 13.481 & 12.854 & 12.717 & $ -31.979$ & 4552 & & & & \\
263 & 276.659752 & $-33.951805$ & 13.489 & 12.884 & 12.727 & $ -5.578$ & 4553 & & & & \\
264 & 276.545369 & $-33.932869$ & 13.493 & 12.880 & 12.781 & $ -61.467$ & 4701 & & & & \\
266 & 276.533760 & $-33.787659$ & 13.505 & 12.874 & 12.744 & $ -68.790$ & 4562 & & & & \\
267 & 276.616984 & $-33.586750$ & 13.506 & 12.954 & 12.792 & $-101.202$ & 4699 & & & & \\
268 & 276.494876 & $-33.647800$ & 13.515 & 12.919 & 12.794 & 71.564 & 4681 & & & & *\\
269 & 276.764459 & $-33.589680$ & 13.518 & 12.940 & 12.784 & 69.289 & 4635 & & & & \\
270 & 276.614926 & $-33.938892$ & 13.526 & 12.948 & 12.808 & 65.396 & 4680 & & & & \\
271 & 276.465499 & $-33.869881$ & 13.526 & 12.857 & 12.692 & 29.413 & 4372 & 2.3 &$+0.48$& 0.6 & \\
272 & 276.827432 & $-33.753338$ & 13.529 & 12.833 & 12.770 & $-104.732$ & 4559 & & & & \\
273 & 276.785331 & $-33.659843$ & 13.538 & 12.873 & 12.710 & 223.568 & 4381 & & & & \\
274 & 276.842104 & $-33.732456$ & 13.541 & 12.852 & 12.714 & $ -21.443$ & 4381 & & & & \\
275 & 276.765440 & $-33.581825$ & 13.554 & 12.999 & 12.817 & 68.660 & 4626 & 2.3 &$-0.76$& 1.0 & \\
276 & 276.728640 & $-33.949757$ & 13.557 & 12.951 & 12.756 & $-121.711$ & 4446 & & & & \\
277 & 276.505794 & $-33.774723$ & 13.561 & 12.941 & 12.782 & $ -51.159$ & 4515 & & & & \\
278 & 276.714377 & $-33.724106$ & 13.581 & 13.007 & 12.860 & 16.094 & 4672 & & & & \\
279 & 276.619384 & $-33.610786$ & 13.591 & 12.968 & 12.842 & 50.290 & 4596 & & & & \\
280 & 276.699742 & $-33.801800$ & 13.600 & 13.064 & 12.896 & $ -22.286$ & 4722 & & & & \\
281 & 276.632739 & $-33.920406$ & 13.601 & 12.931 & 12.843 & $ -74.670$ & 4565 & & & & \\
282 & 276.603736 & $-33.792931$ & 13.606 & 13.010 & 12.886 & $ -59.064$ & 4678 & & & & \\
283 & 276.839527 & $-33.652874$ & 13.614 & 12.951 & 12.909 & $ -47.148$ & 4717 & & & & *\\
284 & 276.724473 & $-33.882050$ & 13.619 & 12.995 & 12.858 & $ -6.884$ & 4555 & & & & \\
285 & 276.810256 & $-33.899017$ & 13.632 & 12.985 & 12.875 & 24.234 & 4562 & & & & *\\
286 & 276.629095 & $-33.587997$ & 13.646 & 13.024 & 12.914 & $-104.407$ & 4645 & & & & \\
287 & 276.860156 & $-33.865948$ & 13.646 & 13.033 & 12.895 & 75.530 & 4578 & & & & \\
288 & 276.589341 & $-33.604183$ & 13.659 & 13.078 & 12.939 & $ -63.114$ & 4682 & & & & \\
291 & 276.787583 & $-33.871380$ & 13.673 & 13.065 & 12.954 & $ -22.181$ & 4673 & & & & \\
292 & 276.501441 & $-33.769341$ & 13.689 & 13.075 & 12.971 & $ -60.436$ & 4688 & & & & \\
294 & 276.807204 & $-33.744389$ & 13.705 & 13.002 & 12.912 & $ -92.853$ & 4469 & & & & \\
295 & 276.671937 & $-33.722317$ & 13.712 & 13.142 & 13.001 & $ -7.833$ & 4704 & & & & \\
296 & 276.472514 & $-33.777775$ & 13.719 & 13.094 & 12.963 & 121.918 & 4578 & & & & \\
297 & 276.686566 & $-33.655930$ & 13.744 & 13.148 & 13.021 & $-110.334$ & 4669 & & & & \\
298 & 276.665050 & $-33.818905$ & 13.746 & 13.132 & 13.029 & 87.153 & 4684 & & & & \\
299 & 276.859868 & $-33.871136$ & 13.748 & 13.085 & 12.995 & $-105.991$ & 4572 & & & & \\
300 & 276.442637 & $-33.752930$ & 13.750 & 13.103 & 13.048 & $ -23.401$ & 4737 & & & & *\\
301 & 276.695536 & $-33.764545$ & 13.751 & 13.160 & 13.029 & $ -68.258$ & 4669 & & & & \\
302 & 276.749584 & $-33.841690$ & 13.752 & 13.130 & 13.059 & $-165.133$ & 4752 & & & & \\
304 & 276.677892 & $-33.890118$ & 13.758 & 13.194 & 12.984 & 54.041 & 4521 & & & & \\
305 & 276.676672 & $-33.567745$ & 13.771 & 13.127 & 13.063 & $ -54.614$ & 4715 & & & & \\
306 & 276.686629 & $-33.707558$ & 13.777 & 13.146 & 13.024 & $ -57.854$ & 4581 & & & & \\
307 & 276.712714 & $-33.944599$ & 13.782 & 13.237 & 13.092 & $ -73.813$ & 4760 & & & & *\\
308 & 276.542295 & $-33.749485$ & 13.790 & 13.240 & 13.092 & $-103.250$ & 4746 & & & & \\
309 & 276.845057 & $-33.643093$ & 13.797 & 13.196 & 13.031 & $ -7.766$ & 4542 & & & & \\
310 & 276.737301 & $-33.623081$ & 13.801 & 13.204 & 13.097 & 68.140 & 4724 & & & & \\
311 & 276.855065 & $-33.892948$ & 13.813 & 13.230 & 13.106 & $ -10.062$ & 4706 & & & & \\
312 & 276.805007 & $-33.749069$ & 13.817 & 13.247 & 13.105 & $ -18.932$ & 4695 & & & & \\
313 & 276.803927 & $-33.834248$ & 13.821 & 13.261 & 13.126 & 29.782 & 4744 & 2.4 &$-0.71$& 0.6 & *\\
314 & 276.622601 & $-33.794971$ & 13.822 & 13.150 & 13.024 & $ -60.428$ & 4460 & & & & \\
315 & 276.821083 & $-33.773975$ & 13.822 & 13.218 & 13.084 & 96.113 & 4618 & & & & *\\
316 & 276.689193 & $-33.710064$ & 13.845 & 13.207 & 13.090 & 0.896 & 4575 & & & & *\\
317 & 276.527656 & $-33.798317$ & 13.846 & 13.287 & 13.138 & $ -20.465$ & 4716 & & & & \\
318 & 276.605035 & $-33.901566$ & 13.846 & 13.198 & 13.094 & $ -8.413$ & 4583 & & & & \\
319 & 276.676819 & $-33.581738$ & 13.852 & 13.225 & 13.053 & $ -58.955$ & 4460 & & & & \\
320 & 276.826131 & $-33.769146$ & 13.853 & 13.137 & 13.051 & $ -41.389$ & 4444 & & & & *\\
321 & 276.547129 & $-33.646622$ & 13.857 & 13.271 & 13.155 & 11.187 & 4736 & & & & \\
322 & 276.625298 & $-33.608067$ & 13.858 & 13.294 & 13.140 & $ -54.255$ & 4687 & & & & \\
323 & 276.850420 & $-33.753113$ & 13.859 & 13.172 & 13.081 & $ -52.017$ & 4508 & & & & *\\
324 & 276.648919 & $-33.838467$ & 13.868 & 13.295 & 13.114 & 29.848 & 4577 & & & & *\\
325 & 276.887809 & $-33.831329$ & 13.871 & 13.297 & 13.117 & $ -84.110$ & 4569 & & & & *\\
326 & 276.520313 & $-33.606983$ & 13.871 & 13.248 & 13.174 & $ -13.705$ & 4753 & & & & \\
327 & 276.743296 & $-33.606194$ & 13.874 & 13.244 & 13.065 & 20.670 & 4431 & 2.5 &$-0.31$& 0.3 & \\
328 & 276.748212 & $-33.826157$ & 13.875 & 13.305 & 13.185 & $ -29.200$ & 4761 & & & & \\
329 & 276.494197 & $-33.811489$ & 13.876 & 13.340 & 13.159 & 10.821 & 4690 & & & & \\
330 & 276.905290 & $-33.690044$ & 13.879 & 13.243 & 13.114 & $ -21.543$ & 4542 & & & & \\
331 & 276.661029 & $-33.805126$ & 13.882 & 13.245 & 13.169 & $ -85.801$ & 4696 & & & & \\
332 & 276.725178 & $-33.922646$ & 13.894 & 13.301 & 13.118 & 74.659 & 4514 & & & & \\
333 & 276.803651 & $-33.785221$ & 13.895 & 13.274 & 13.161 & $-160.010$ & 4630 & & & & \\
334 & 276.573211 & $-33.613113$ & 13.897 & 13.384 & 13.189 & 30.030 & 4718 & & & & *\\
335 & 276.754422 & $-33.812862$ & 13.899 & 13.284 & 13.177 & $ -3.150$ & 4666 & & & & \\
336 & 276.775634 & $-33.868172$ & 13.902 & 13.205 & 13.179 & $ -19.149$ & 4661 & & & & *\\
337 & 276.609719 & $-33.899956$ & 13.910 & 13.284 & 13.186 & 86.522 & 4663 & & & & \\
338 & 276.802770 & $-33.607288$ & 13.910 & 13.270 & 13.182 & $ -31.855$ & 4651 & & & & *\\
339 & 276.599163 & $-33.751358$ & 13.916 & 13.369 & 13.197 & 21.543 & 4682 & & & & \\
340 & 276.862496 & $-33.881008$ & 13.916 & 13.287 & 13.203 & $ -7.419$ & 4688 & & & & \\
342 & 276.619675 & $-33.905045$ & 13.925 & 13.343 & 13.216 & 123.234 & 4708 & & & & \\
343 & 276.694856 & $-33.562214$ & 13.926 & 13.257 & 13.149 & 69.723 & 4519 & & & & \\
344 & 276.586781 & $-33.832466$ & 13.931 & 13.301 & 13.158 & 214.266 & 4528 & & & & \\
347 & 276.736184 & $-33.875690$ & 13.949 & 13.316 & 13.213 & $ -45.041$ & 4624 & & & & \\
348 & 276.578552 & $-33.619892$ & 13.950 & 13.352 & 13.202 & 5.935 & 4600 & & & & \\
349 & 276.737919 & $-33.695145$ & 13.956 & 13.350 & 13.207 & $ -65.841$ & 4590 & & & & \\
350 & 276.456610 & $-33.796585$ & 13.957 & 13.347 & 13.225 & $-120.618$ & 4647 & & & & \\
351 & 276.564966 & $-33.635128$ & 13.960 & 13.423 & 13.248 & 128.623 & 4706 & 2.5 &$+0.17$& 1.2 & \\
353 & 276.861593 & $-33.801018$ & 13.974 & 13.438 & 13.274 & 22.706 & 4728 & & & & \\
354 & 276.722013 & $-33.592499$ & 13.974 & 13.376 & 13.238 & 46.922 & 4630 & & & & \\
355 & 276.749304 & $-33.665966$ & 13.975 & 13.357 & 13.210 & $ -58.417$ & 4547 & & & & \\
356 & 276.762247 & $-33.786091$ & 13.982 & 13.460 & 13.272 & 20.393 & 4702 & & & & \\
357 & 276.785991 & $-33.896404$ & 13.985 & 13.341 & 13.224 & $ -84.393$ & 4553 & & & & \\
358 & 276.833052 & $-33.776608$ & 13.989 & 13.368 & 13.235 & $ -0.947$ & 4572 & & & & \\
359 & 276.433122 & $-33.740566$ & 13.989 & 13.411 & 13.292 & $ -32.267$ & 4753 & & & & \\
360 & 276.691675 & $-33.846764$ & 13.993 & 13.347 & 13.275 & $ -38.991$ & 4679 & & & & \\
361 & 276.705482 & $-33.740154$ & 13.996 & 13.411 & 13.280 & 37.993 & 4687 & & & & \\
362 & 276.617111 & $-33.760021$ & 13.996 & 13.410 & 13.278 & 7.110 & 4684 & & & & \\
363 & 276.493794 & $-33.799362$ & 13.996 & 13.417 & 13.299 & 30.066 & 4750 & & & & \\
364 & 276.697108 & $-33.596195$ & 13.996 & 13.460 & 13.297 & 49.761 & 4741 & & & & \\
365 & 276.575708 & $-33.577957$ & 14.000 & 13.428 & 13.271 & 0.930 & 4656 & & & & \\
366 & 276.663677 & $-33.647522$ & 14.002 & 13.400 & 13.306 & 12.679 & 4750 & & & & \\
367 & 276.779362 & $-33.567799$ & 14.002 & 13.382 & 13.217 & 188.004 & 4495 & & & & \\
368 & 276.871279 & $-33.792786$ & 14.006 & 13.428 & 13.318 & $ -35.508$ & 4764 & & & & \\
369 & 276.580553 & $-33.632629$ & 14.007 & 13.438 & 13.271 & $ -80.578$ & 4634 & & & & \\
370 & 276.711834 & $-33.703365$ & 14.013 & 13.394 & 13.306 & 48.722 & 4714 & & & & \\
371 & 276.743577 & $-33.898552$ & 14.014 & 13.444 & 13.263 & $ -93.685$ & 4581 & & & & *\\
373 & 276.441719 & $-33.814545$ & 14.024 & 13.424 & 13.276 & 2.274 & 4601 & & & & \\
374 & 276.570897 & $-33.627617$ & 14.025 & 13.431 & 13.228 & $ -16.193$ & 4468 & & & & \\
376 & 276.747908 & $-33.773846$ & 14.027 & 13.360 & 13.322 & $ -31.917$ & 4718 & & & & \\
377 & 276.641959 & $-33.642696$ & 14.027 & 13.439 & 13.293 & 9.039 & 4638 & & & & *\\
378 & 276.686341 & $-33.926971$ & 14.032 & 13.484 & 13.341 & 119.348 & 4758 & & & & \\
379 & 276.867112 & $-33.799599$ & 14.035 & 13.406 & 13.344 & 36.952 & 4755 & & & & \\
380 & 276.797940 & $-33.867813$ & 14.041 & 13.329 & 13.276 & 10.755 & 4542 & & & & \\
381 & 276.615896 & $-33.679054$ & 14.044 & 13.455 & 13.337 & $ -41.469$ & 4718 & & & & \\
382 & 276.642599 & $-33.707905$ & 14.051 & 13.414 & 13.319 & 29.732 & 4642 & & & & *\\
383 & 276.552809 & $-33.793800$ & 14.052 & 13.441 & 13.273 & 58.636 & 4514 & & & & \\
384 & 276.767664 & $-33.790257$ & 14.054 & 13.476 & 13.288 & $ -40.420$ & 4542 & & & & \\
385 & 276.519644 & $-33.920509$ & 14.060 & 13.538 & 13.278 & $ -29.684$ & 4504 & & & & *\\
386 & 276.778487 & $-33.904800$ & 14.063 & 13.482 & 13.386 & 7.124 & 4799 & & & & \\
387 & 276.738181 & $-33.939156$ & 14.068 & 13.462 & 13.390 & 93.169 & 4797 & & & & *\\
388 & 276.529173 & $-33.642719$ & 14.071 & 13.505 & 13.331 & $-107.163$ & 4624 & & & & \\
389 & 276.805911 & $-33.751480$ & 14.078 & 13.399 & 13.380 & 179.763 & 4737 & & & & *\\
390 & 276.654480 & $-33.594334$ & 14.081 & 13.422 & 13.341 & $ -8.853$ & 4621 & & & & \\
392 & 276.628139 & $-33.834789$ & 14.097 & 13.540 & 13.394 & 7.430 & 4726 & & & & *\\
393 & 276.710695 & $-33.846008$ & 14.100 & 13.400 & 13.312 & 84.567 & 4484 & & & & \\
394 & 276.689541 & $-33.778378$ & 14.107 & 13.531 & 13.367 & 28.170 & 4616 & & & & \\
395 & 276.514699 & $-33.757675$ & 14.110 & 13.464 & 13.352 & 143.624 & 4572 & & & & \\
396 & 276.694978 & $-33.943527$ & 14.117 & 13.546 & 13.403 & 90.426 & 4689 & & & & \\
397 & 276.828208 & $-33.699005$ & 14.118 & 13.473 & 13.418 & 31.582 & 4731 & & & & \\
398 & 276.515313 & $-33.784389$ & 14.119 & 13.550 & 13.332 & $ -78.551$ & 4493 & & & & \\
399 & 276.634575 & $-33.901833$ & 14.126 & 13.579 & 13.433 & $ -52.671$ & 4755 & & & & \\
400 & 276.829987 & $-33.806454$ & 14.127 & 13.557 & 13.385 & $ -47.000$ & 4605 & & & & \\
401 & 276.744083 & $-33.730270$ & 14.129 & 13.516 & 13.358 & $ -64.671$ & 4530 & & & & \\
402 & 276.879971 & $-33.675831$ & 14.138 & 13.436 & 13.357 & 13.679 & 4500 & & & & \\
403 & 276.873784 & $-33.655762$ & 14.143 & 13.553 & 13.461 & 181.139 & 4785 & & & & \\
404 & 276.750878 & $-33.678814$ & 14.145 & 13.559 & 13.397 & $ -16.703$ & 4593 & & & & \\
406 & 276.773448 & $-33.816742$ & 14.159 & 13.602 & 13.453 & 89.280 & 4713 & & & & \\
407 & 276.538950 & $-33.857552$ & 14.167 & 13.626 & 13.426 & $ -59.382$ & 4617 & & & & \\
408 & 276.702749 & $-33.817944$ & 14.168 & 13.557 & 13.489 & $ -91.236$ & 4797 & & & & \\
409 & 276.578957 & $-33.705822$ & 14.171 & 13.622 & 13.446 & 0.475 & 4665 & & & & \\
410 & 276.753490 & $-33.738415$ & 14.173 & 13.552 & 13.448 & 24.100 & 4659 & & & & \\
411 & 276.836391 & $-33.686768$ & 14.174 & 13.593 & 13.453 & 21.788 & 4669 & & & & \\
412 & 276.501238 & $-33.637436$ & 14.174 & 13.595 & 13.412 & $-161.418$ & 4563 & & & & \\
413 & 276.662487 & $-33.626492$ & 14.179 & 13.651 & 13.492 & 50.701 & 4778 & & & & \\
414 & 276.628673 & $-33.566936$ & 14.185 & 13.516 & 13.421 & $ -5.715$ & 4555 & & & & \\
415 & 276.534514 & $-33.915627$ & 14.195 & 13.583 & 13.469 & 58.092 & 4660 & & & & \\
416 & 276.673953 & $-33.712120$ & 14.197 & 13.610 & 13.502 & $-111.533$ & 4751 & & & & \\
417 & 276.561145 & $-33.898586$ & 14.197 & 13.540 & 13.448 & 47.080 & 4592 & & & & \\
419 & 276.578991 & $-33.573586$ & 14.206 & 13.587 & 13.472 & $ -41.636$ & 4641 & & & & \\
420 & 276.687936 & $-33.576183$ & 14.212 & 13.551 & 13.441 & $ -69.762$ & 4535 & 2.5 &$+0.05$& 0.5 & *\\
422 & 276.852338 & $-33.874542$ & 14.215 & 13.578 & 13.489 & $ -20.986$ & 4650 & & & & \\
423 & 276.852706 & $-33.851723$ & 14.232 & 13.667 & 13.550 & 45.436 & 4782 & & & & \\
424 & 276.886130 & $-33.787621$ & 14.234 & 13.670 & 13.553 & 101.421 & 4785 & & & & \\
425 & 276.661265 & $-33.711018$ & 14.238 & 13.657 & 13.487 & 23.665 & 4587 & & & & \\
426 & 276.616342 & $-33.801567$ & 14.238 & 13.734 & 13.527 & $-107.099$ & 4704 & & & & \\
427 & 276.858769 & $-33.890724$ & 14.238 & 13.579 & 13.486 & $-152.559$ & 4575 & & & & \\
428 & 276.683750 & $-33.836472$ & 14.242 & 13.685 & 13.528 & $ -70.446$ & 4692 & & & & \\
429 & 276.794492 & $-33.592255$ & 14.245 & 13.644 & 13.565 & 3.931 & 4796 & & & & \\
430 & 276.767692 & $-33.907009$ & 14.247 & 13.673 & 13.511 & $-154.454$ & 4622 & & & & \\
431 & 276.561013 & $-33.621979$ & 14.248 & 13.682 & 13.531 & 82.186 & 4692 & & & & *\\
433 & 276.801658 & $-33.837830$ & 14.256 & 13.632 & 13.536 & $ -60.959$ & 4670 & & & & *\\
434 & 276.665024 & $-33.724117$ & 14.258 & 13.665 & 13.550 & $ -22.264$ & 4713 & & & & *\\
435 & 276.768679 & $-33.857334$ & 14.259 & 13.664 & 13.559 & 99.095 & 4730 & & & & \\
436 & 276.493553 & $-33.646214$ & 14.263 & 13.713 & 13.588 & $ -50.196$ & 4822 & & & & *\\
437 & 276.827726 & $-33.647717$ & 14.268 & 13.643 & 13.542 & $ -52.388$ & 4655 & & & & \\
438 & 276.594804 & $-33.845116$ & 14.271 & 13.742 & 13.515 & 5.719 & 4573 & & & & \\
440 & 276.631808 & $-33.564571$ & 14.278 & 13.711 & 13.603 & 14.329 & 4819 & & & & \\
441 & 276.700830 & $-33.637337$ & 14.281 & 13.799 & 13.534 & $ -18.383$ & 4598 & & & & \\
443 & 276.727795 & $-33.730831$ & 14.297 & 13.769 & 13.573 & $ -32.104$ & 4663 & & & & \\
445 & 276.472846 & $-33.808666$ & 14.310 & 13.670 & 13.574 & 73.543 & 4634 & 2.6 &$+0.25$& 0.7 & \\
446 & 276.438307 & $-33.774590$ & 14.311 & 13.721 & 13.542 & $ -8.678$ & 4544 & & & & \\
447 & 276.777476 & $-33.775871$ & 14.316 & 13.739 & 13.634 & $ -81.331$ & 4786 & & & & \\
448 & 276.706230 & $-33.843826$ & 14.316 & 13.735 & 13.588 & $ -20.233$ & 4649 & & & & \\
449 & 276.798371 & $-33.735714$ & 14.320 & 13.855 & 13.631 & 5.772 & 4765 & & & & \\
451 & 276.887413 & $-33.852146$ & 14.324 & 13.738 & 13.636 & $-184.313$ & 4762 & 2.6 &$-1.05$& 0.8 & \\
452 & 276.661034 & $-33.891323$ & 14.334 & 13.821 & 13.654 & $ -76.626$ & 4794 & 2.6 &$-0.99$& 0.7 & \\
453 & 276.690616 & $-33.808006$ & 14.336 & 13.754 & 13.596 & $ -48.148$ & 4615 & & & & \\
454 & 276.698507 & $-33.939789$ & 14.343 & 13.800 & 13.657 & 115.364 & 4773 & & & & \\
455 & 276.845583 & $-33.661392$ & 14.344 & 13.746 & 13.667 & 65.842 & 4802 & & & & \\
456 & 276.593340 & $-33.672127$ & 14.349 & 13.759 & 13.642 & $ -23.056$ & 4719 & & & & \\
457 & 276.833423 & $-33.746063$ & 14.353 & 13.767 & 13.581 & $ -23.375$ & 4524 & & & & \\
459 & 276.615328 & $-33.576347$ & 14.368 & 13.740 & 13.606 & 49.302 & 4561 & & & & \\
460 & 276.722855 & $-33.943268$ & 14.391 & 13.838 & 13.708 & $ -73.249$ & 4781 & & & & \\
461 & 276.834942 & $-33.635342$ & 14.396 & 13.863 & 13.666 & $ -19.355$ & 4643 & 2.7 &$-1.50$& 0.8 & \\
462 & 276.789630 & $-33.841022$ & 14.400 & 13.855 & 13.727 & 4.985 & 4813 & & & & \\
463 & 276.467867 & $-33.749195$ & 14.400 & 13.783 & 13.650 & $-172.922$ & 4595 & & & & \\
464 & 276.657623 & $-33.651539$ & 14.406 & 13.786 & 13.632 & $ -66.065$ & 4526 & & & & \\
465 & 276.631871 & $-33.704411$ & 14.417 & 13.879 & 13.739 & $-121.778$ & 4806 & & & & \\
466 & 276.861515 & $-33.729511$ & 14.417 & 13.903 & 13.742 & $ -59.983$ & 4806 & & & & \\
467 & 276.742712 & $-33.819698$ & 14.422 & 13.889 & 13.745 & $ -9.504$ & 4802 & & & & \\
468 & 276.511268 & $-33.792122$ & 14.422 & 13.791 & 13.707 & 56.092 & 4696 & & & & \\
469 & 276.581345 & $-33.660347$ & 14.425 & 13.834 & 13.711 & $ -44.438$ & 4699 & & & & \\
471 & 276.439088 & $-33.761520$ & 14.447 & 13.839 & 13.738 & 35.018 & 4717 & & & & \\
472 & 276.823834 & $-33.629124$ & 14.464 & 13.925 & 13.731 & 131.626 & 4635 & & & & \\
473 & 276.670917 & $-33.612740$ & 14.466 & 13.942 & 13.796 & 24.364 & 4833 & & & & *\\
474 & 276.547004 & $-33.669544$ & 14.475 & 13.910 & 13.772 & $ -17.630$ & 4733 & & & & \\
476 & 276.661846 & $-33.655987$ & 14.491 & 13.900 & 13.829 & $ -29.627$ & 4859 & & & & \\
477 & 276.853016 & $-33.748569$ & 14.496 & 13.867 & 13.809 & $ -75.148$ & 4768 & 2.7 &$-1.09$& 0.7 & \\
478 & 276.533779 & $-33.846905$ & 14.501 & 13.982 & 13.812 & $ -38.002$ & 4772 & & & & \\
481 & 276.777363 & $-33.671219$ & 14.502 & 13.896 & 13.736 & $ -11.567$ & 4543 & & & & *\\
482 & 276.573957 & $-33.858578$ & 14.505 & 14.030 & 13.800 & $-154.032$ & 4722 & 2.7 &$-0.78$& 0.9 & \\
483 & 276.603613 & $-33.823669$ & 14.506 & 13.932 & 13.766 & 31.359 & 4618 & & & & \\
484 & 276.594676 & $-33.571365$ & 14.507 & 13.849 & 13.829 & $ -6.386$ & 4810 & & & & *\\
485 & 276.530138 & $-33.766701$ & 14.511 & 13.876 & 13.792 & 129.335 & 4684 & & & & \\
486 & 276.865712 & $-33.679306$ & 14.515 & 13.996 & 13.838 & $ -56.181$ & 4801 & & & & \\
487 & 276.849420 & $-33.771736$ & 14.519 & 13.847 & 13.786 & $-141.982$ & 4631 & & & & \\
488 & 276.598639 & $-33.744186$ & 14.521 & 13.967 & 13.832 & $ -3.129$ & 4771 & & & & \\
489 & 276.875366 & $-33.742172$ & 14.521 & 13.810 & 13.769 & 73.034 & 4577 & & & & \\
490 & 276.712242 & $-33.794239$ & 14.531 & 13.928 & 13.856 & $ -2.514$ & 4810 & & & & \\
491 & 276.751951 & $-33.710484$ & 14.534 & 13.883 & 13.786 & $ -47.811$ & 4593 & & & & \\
493 & 276.631945 & $-33.662464$ & 14.538 & 13.956 & 13.862 & 54.638 & 4813 & 2.7 &$-1.12$& 0.9 & *\\
494 & 276.572343 & $-33.570671$ & 14.542 & 13.939 & 13.812 & $ -28.906$ & 4653 & & & & \\
495 & 276.548845 & $-33.833427$ & 14.549 & 13.924 & 13.793 & 34.963 & 4575 & & & & \\
496 & 276.765063 & $-33.868202$ & 14.555 & 13.990 & 13.859 & 77.859 & 4742 & & & & \\
497 & 276.830924 & $-33.896267$ & 14.555 & 13.988 & 13.816 & $ -30.904$ & 4612 & & & & \\
498 & 276.771491 & $-33.669971$ & 14.559 & 13.951 & 13.817 & $ -23.933$ & 4610 & & & & \\
499 & 276.633311 & $-33.762764$ & 14.567 & 13.899 & 13.838 & 31.087 & 4650 & & & & \\
500 & 276.602649 & $-33.731819$ & 14.567 & 14.041 & 13.883 & $ -1.571$ & 4787 & & & & \\
501 & 276.858813 & $-33.772926$ & 14.568 & 13.883 & 13.878 & 174.776 & 4759 & & & & \\
503 & 276.436893 & $-33.808731$ & 14.580 & 13.968 & 13.852 & $ -36.423$ & 4659 & & & & \\
504 & 276.481768 & $-33.625542$ & 14.581 & 13.971 & 13.872 & $ -67.245$ & 4718 & & & & \\
505 & 276.537031 & $-33.778503$ & 14.584 & 14.036 & 13.893 & $ -9.908$ & 4767 & & & & \\
506 & 276.502711 & $-33.686172$ & 14.585 & 13.967 & 13.838 & $ -10.462$ & 4604 & & & & \\
507 & 276.736185 & $-33.691521$ & 14.591 & 13.961 & 13.833 & $ -87.751$ & 4566 & & & & \\
508 & 276.761504 & $-33.739830$ & 14.594 & 13.971 & 13.862 & $-231.533$ & 4638 & 2.7 &$-0.71$& 0.9 & \\
509 & 276.856836 & $-33.670521$ & 14.596 & 13.998 & 13.898 & $-150.306$ & 4737 & 2.7 &$-0.76$& 0.7 & \\
511 & 276.586662 & $-33.710934$ & 14.606 & 13.974 & 13.869 & 106.628 & 4629 & & & & \\
512 & 276.620755 & $-33.646839$ & 14.620 & 14.124 & 13.932 & $ -19.865$ & 4776 & & & & \\
513 & 276.868502 & $-33.816536$ & 14.620 & 14.136 & 13.959 & 39.173 & 4851 & & & & \\
514 & 276.727030 & $-33.799290$ & 14.624 & 14.014 & 13.922 & 42.665 & 4727 & & & & *\\
\end{longtable}
\tablefoot{Meaning of the columns: (1): Target number in this programme; (2):
Right ascension from 2MASS; (3): Declination from 2MASS; (4), (5), and (6):
2MASS $J$-, $H$-, and $K$-magnitude; (7): Heliocentric radial velocity; (8):
Effective temperature; (9): Surface gravity $\log g$; (10): Metallicity; (11):
Li abundance; (12): Flag for foreground stars.}
\label{tabappend}
\section{Discussion}
\subsection{The frequency of Li rich stars}
The three most Li-rich stars in our sample are located on the upper RGB,
clearly above the bump and the two RCs. The faintest of our Li-rich stars
(\#123), at a $K$-magnitude of 11\fm61, is more than one magnitude brighter
than the bright RC (peak at $K\sim12\fm65$). Taking into account the
$K$-magnitude spread of the RCs of 0\fm22 (David Nataf, private communication),
this means that the faintest Li-rich star is brighter than the bright RC peak
by 4.7 standard deviations. Hence, even considering the error in distance caused
by the spread within the Bulge, it is extremely unlikely that the Li-rich stars
are related to either the RGB bump or clump. On the other hand, the
Li-{\em detected} stars are scattered all along the RGB, from below the bump to
the tip. Neither at the bump nor at the RCs do we see an overdensity of
Li-detected stars. In contrast, the {\em fraction} of Li-detected stars on the
upper RGB ($J_0 < 12\fm0$) is 18.8\%, much higher than below this limit
(5.4\%). This important result of the present study is illustrated in
Fig.~\ref{Lifract}, which shows a histogram of the fraction of Li-detected
stars as a function of $J_0$. The fraction of Li-detected stars also increases
as a function of $(J-K_s)_0$, with a local maximum of 26.3\% between
$(J-K_s)_0=0.8$ and 0.9.
\begin{figure}
\centering
\includegraphics[width=\linewidth,bb=83 366 540 699,clip]{Lifract.pdf}
\caption{Fraction of Li-detected stars as a function of $J_0$.}
\label{Lifract}
\end{figure}
Recently, U07 identified four AGB stars in the PG3 field with Li abundance
of $\log\epsilon({\rm Li})=0.8$, 0.8, 1.1, and 2.0, among a sample of 27
long-period AGB variables. The fraction of {\em upper} RGB stars with
detectable Li line is similar to that found among AGB stars, accordingly it is
possible that the Li-rich AGB stars inherited their Li from the preceding RGB
phase. Indeed, the three Li-rich stars identified in the present study could be
early AGB stars instead of RGB stars because they are brighter than the RC.
Because it is impossible to separate RGB and early AGB stars by means of our
photometric and spectroscopic data alone, asteroseismological methods would be
needed to better define their precise evolutionary state.
\subsection{Mass loss from Li-rich stars?}
It has been speculated that the Li enrichment in K-type giants is accompanied
by a mass-loss episode \citep{dlR96,dlR97}, which has, however, later been
questioned \citep{FW98,Jas99}. To investigate if there is enhanced
{\em dust} mass-loss from our target stars, we cross-identified them with
sources detected by the WISE space observatory \citep{Wri10}. Within a search
radius of 1\farcs2 we found counterparts for 333 of our targets in the WISE
catalogue\footnote{http://irsa.ipac.caltech.edu/}. We inspected $J-K$ vs.\
$K - \rm{WISE}~[3.4{\rm \mu m}]$, $[4.6{\rm \mu m}]$, and $[12{\rm \mu m}]$
colour-colour diagrams of these stars. The 22\,${\rm \mu m}$ WISE band proved to
have too low a sensitivity to reliable detect more than just the very brightest
of our targets. A version of the $J-K$ vs.\ $K - \rm{WISE}~[12{\rm \mu m}]$
colour-colour diagram is displayed in Fig.~\ref{WISE_CCD}. The x-axis range was
restricted to the reddest part of the sample, because at the bluer (and
hence fainter) end of the sample the noise level is very high. The few red
outliers in the colour-colour diagrams were identified in the 2MASS images to
be visual pairs or triples that are resolved by 2MASS, but are probably
unresolved in the WISE observations. The three Li-rich stars have very
inconspicuous $K - \rm{WISE}$ colours, strongly suggesting that they do not
suffer enhanced dusty mass-loss. For comparison, the Miras investigated by U07,
which suffer a total mass-loss rate of $10^{-8} M_{\odot} {\rm yr}^{-1}$ or more,
have $K - [12{\rm \mu m}]$ colours of 1.17 or redder.
\begin{figure}
\centering
\includegraphics[width=\linewidth,bb=71 373 538 699,clip]{JK_vs_K-WISE12.pdf}
\caption{$K - \rm{WISE} [12{\rm \mu m}]$ vs.\ $J-K$ colour-colour diagram of
our sample stars. The grey diamond symbols represent the three Li-rich
stars.}
\label{WISE_CCD}
\end{figure}
To search also for enhanced {\em gas} mass-loss, we inspected the spectra of
the Li-rich stars around the H$\alpha$ line in comparison with otherwise
similar Li-poor stars, namely \#051 and \#131. Gas mass-loss would be
detectable through blue-shifted asymmetries in the H$\alpha$ line profile, see
e.g.\ \citet{Mes09} for a sensitive search for gas mass-loss in red giants. The
most Li-rich sample star (\#042) has a slightly broader H$\alpha$ profile than
the comparison stars, but no asymmetry. The other two Li-rich stars have a
symmetric profile as well. We conclude from this that within the limits of
resolution and S/N of our spectra, the Li-rich stars do not suffer enhanced
loss of either dust nor gas, confirming the conclusions of \citet{FW98} and
\citet{Jas99}.
\subsection{A trend of Li abundance with temperature}
Recently, a study of Li-rich giants in the GB was presented by \citet{Gon09}.
Among a sample of 417 stars, chosen from a narrow brightness range located
about $0\fm7$ above the expected RC, they found 13 stars with detectable Li
line. \citet{Gon09} concluded that it is unlikely that the Li-rich stars are
connected to either the RGB bump or the RC if they are genuine GB stars at a
mean distance of 8\,kpc. Although the authors' sample covers only a narrow
luminosity range along the RGB, their results agree with ours. \citet{Gon09}
report a decrease of the Li abundance in the Li-detected stars with decreasing
temperature in their sample, which again agrees with our result. A comparison
with their LTE results is presented in Fig.~\ref{ALiTeff}. Our sample extends
the trend found by \citet{Gon09} to lower temperatures. Just as in
\citet{Gon09}, our Li-rich stars also deviate from this trend, and fall in
a similar temperature range at $T_{\rm eff} = \sim 4100 - 4300$\,K. One
difference to our study is, however, that \citet{Gon09} find significantly
higher Li-abundances at the hot end of their sample ($T_{\rm eff} \geq 4600$\,K)
than we do. This difference is mostly caused by four stars in the \citet{Gon09}
sample at $T_{\rm eff} \geq 4600$\,K and $\log\epsilon({\rm Li})\geq1.5$. Those
stars also have a lower surface gravity ($\log g = 1.9 - 2.1$) than we estimate
for our sample stars in this temperature range ($\log g = 2.3 - 2.7$). Thus,
there could be a difference in evolutionary state between these groups of
stars, also in terms of Li abundance, but we are not sure if this fully
accounts for the found difference.
\begin{figure}
\centering
\includegraphics[width=\linewidth,bb=84 367 536 698,clip]{ALi_vs_Teff.pdf}
\caption{Lithium abundance as a function of effective temperatures of RGB
stars. The data points from our study are plotted as plus symbols, while
the open triangles represent the LTE measurements from \citet{Gon09}.}
\label{ALiTeff}
\end{figure}
Furthermore, \citet{Gon09} also carefully examined whether the Li-rich stars are
special in any other way, but did not find a difference to the Li-normal stars.
In a similar attempt, we inspected the spectra of the Li-rich and Li-detected
stars in our sample in comparison with Li-poor stars to search for
peculiarities, but we did not find any. We only found two sample stars with
increased strength in the BaII 649.9\,nm and the LaII 652.9\,nm lines. They
probably belong to the so-called barium star class, collecting members of
binary systems that owe their enhanced s-element abundances to a mass-transfer
episode from a more massive AGB companion. These will be discussed in more
detail in a forthcoming paper. The Li-rich and Li-detected stars are perfectly
normal in their BaII and LaII line strengths, which certifies that their Li
overabundance is not the result of mass transfer of Li-enriched matter from a
binary (former AGB) companion. Finally, \citet{Gon09} also conclude that their
Li-rich stars do not suffer enhanced mass-loss, based 2MASS $JHK$ photometry.
The three stars with the highest Li abundance in our sample fall nicely within
the sequence of Li-rich thick disk giants identified by
\citet[][their Fig.~4]{Mon11}. Recently, \citet{Kumar11} also identified
Li-rich giants in the Galactic disk and ascribed them to the red clump, i.e.\
these are stars after the helium core flash. We found no accumulation of
Li-rich stars close to the RCs in our sample. This could be related to a
difference in mass between the stars in our study and that of \citet{Kumar11}:
While there are mostly low-mass stars with a narrow distribution around
$1 M_{\sun}$ in our sample, the disk sample of \citet{Kumar11} likely also
contains stars of higher mass, whose Li abundance may evolve differently. The
thick disk sample of \citet{Mon11} likely contains mostly stars of low mass as
well, although the authors identify one star that could have a mass of up to
4\,$M_{\sun}$, albeit with large uncertainty. The agreement between our results
and those of \citet{Mon11} is probably related to a similar mass spectrum of
the samples.
\section{Conclusions}
There are now several studies that show that at low to intermediate masses
Li-rich stars can be found all along the RGB, not necessarily connected to a
particular phase of the giant branch evolution \citep{Gon09,Mon11,Alc11,Ruch11}.
This means that either the Li enrichment process is stochastic in nature and can
happen at any time on the RGB, or that Li enrichment starts at some point on
the lower RGB and may take a long time to make Li-rich K giants. The activation
of mixing events of variable duration and transport rates below the convective
envelope is conceivable for the Li enrichment \citep{Palm11}. The lack of
a well-defined Li-rich phase on the RGB, as now shown by several recent
studies, contradicts predictions by CB00, who propose that a Li-rich
phase should occur at the bump of the RGB in low-mass stars, such as those that
we have in our sample. The existence of such a Li-rich phase has to be
questioned.
\begin{acknowledgements}
TL acknowledges support from the Austrian Science Fund (FWF) under projects
P~21988-N16 and P~23737-N16, and SU under project P~22911-N16. BA thanks for
the support from contract ASI-INAF I/009/10/0. This publication makes use of
data products from the Two Micron All Sky Survey, which is a joint project of
the University of Massachusetts and the Infrared Processing and Analysis
Center/California Institute of Technology, funded by the National Aeronautics
and Space Administration and the National Science Foundation. This
publication makes use of data products from the Wide-field Infrared Survey
Explorer, which is a joint project of the University of California, Los
Angeles, and the Jet Propulsion Laboratory/California Institute of
Technology, funded by the National Aeronautics and Space Administration.
\end{acknowledgements}
|
\section{Introduction}
We live in an expanding universe. There is very good evidence that this expansion is in fact accelerating \cite{Riess:1998cb, Perlmutter:1998np}, and the central idea of the theory of inflation \cite{Guth:1980zm, Linde:1981mu, Albrecht:1982wi} is that in the distant past, for a period of time not long after the big bang, this behavior was a dominant feature of our universe.
During such periods of accelerated expansion, different regions of the universe lose causal contact with each other. A distant star we see in the sky one night might not be visible anymore the following night, not because it had stopped shining, but simply because the light it emits will never reach us.
On the other hand, we are used to thinking of physics in terms of local quantities and fields, which have correlators that fall off rapidly with distance. Let us consider an inflating universe endowed with a scalar field living in it. If this field is massless, its expectation value may vary widely from place to place.\footnote{Needless to say, there are no such massless scalars with significant couplings to the standard model particles in our universe today, else we would have noticed their profound consequences for low-energy physics (though almost massless scalar fields do occur e.g.~in models of slow roll inflation). Thus we should perhaps not think of this field as matter, but rather as an abstract probe of the properties of de Sitter space itself.}
Clearly, the field locally wants to minimize its energy density by not allowing steep gradients, but even small gradients can lead to vastly different expectation values over large distances. Once a neighboring region disappears behind our cosmological horizon, and we can no longer communicate with it, what is to stop the field in this region from drifting towards an expectation value completely different from what we measure in our vicinity?
Thus small quantum fluctuations may be amplified into a wildly varying field profile.
One might argue that a local observer should not care about expectation values in regions beyond his horizon.
But now let's take a more global point of view. Imagine an observer that looks back at these disconnected regions after inflation has ended (when formerly out of contact regions are becoming accessible again) and samples the average expectation values of the quantum field in each of those regions. What is the distribution of these coarse-grained expectation values (and conjugate momenta) that such an observer would measure?
Questions of this nature have been studied for a long time in the context of cosmological perturbation theory and structure formation \cite{Bardeen:1980kt, Mukhanov:1990me}. There is a large body of work seeking to predict what spectrum of matter density fluctuations a post-inflation observer should observe in the sky, assuming they arise from primordial quantum fluctuations and are modified according to the details of one's favorite inflationary model.
But here we would like to take a further step back, and consider the statistical properties of these coarse-grained variables in phase space.
Given the thermal nature of the de Sitter vacuum, is there ergodicity breaking? Does this system exhibit a glassy structure, as suggested in \cite{AD}?
We would like to shed some light on the global structure of the reduced phase space of expectation values (and conjugate momenta) corresponding to causally disconnected regions on a constant time slice. To this end, we will make use of tools and techniques from spin glass theory (see e.g.~\cite{MPV,CG,DG}), in particular the overlap and triple overlap distributions, which can be understood as order parameters for a non-trivial structure in the space of thermodynamic states \cite{Parisi:1983dx}.
The application of spin glass methods to de Sitter dynamics was pioneered by F.~Denef and D.~Anninos \cite{AD, Denef:2011ee, APhD, Unpub}, who conceived the central ideas and were the first to investigate overlaps of various types on de Sitter spaces of different dimensionalities. They developed the formalism of reduced density matrix based overlaps that we will employ, and also wrote down the simple model that we will study, as an ideal playground to both test these techniques and attempt to tackle questions about the state space of de Sitter. In Section \ref{RevSet} we will review the relevant parts of their work, and remind the reader of some simple definitions and statistical notions borrowed from the theory of spin glasses.
Section \ref{Low} describes a direct attempt to characterize the overlap, which will give us the first few moments of the distribution, while in Section \ref{Graphs} we introduce a slightly more sophisticated method, based on graphs and combinatorics, that allows us to find the complete distribution (at leading order in a late-time expansion). In Section \ref{Triple} we generalize this method to compute the triple overlap distribution, before concluding with a brief discussion of our results.
\section{Setup and Review} \label{RevSet}
\subsection{A Simple Model}
The model we will consider \cite{Unpub} could hardly be any simpler. As a background will take a 1+1 dimensional de Sitter spacetime, whose metric we can write in global coordinates as
\[
\mathrm{d} s^2 = - \mathrm{d} \tau^2 + \cosh^2 \tau\, \mathrm{d} \theta^2 \ ,
\]
where we have set the cosmological constant to unity.
A fixed time $\tau=$ const.~slice is then simply a circle parametrized by the angular variable $\theta \in (0,2\pi)$.
On this background we study the evolution of a real, massless scalar field $\phi(\tau,\theta)$. For a constant time slice we can Fourier decompose its fluctuations in the obvious fashion\footnote{We have chosen a closed slicing here in order to obtain discrete modes.}
\[
\phi = \sum_{n=-\infty}^\infty \phi_n\, e^{i n \theta} \ ,
\]
where reality requires that $\phi_{-n} = \phi_n^\ast$. The Euclidean vacuum wavefunctional \cite{HartHawk, Allen:1985ux} can then be written as
\[\label{EucVac}
\Psi = \prod_{n=1}^\infty \sqrt{2 n \over \pi}\, \exp\left(- n |\phi_n|^2 \right) \ ,
\]
which simply takes the form of an infinite sequence of harmonic oscillators (and we have removed the zero mode).\footnote{This is closely related to the string world-sheet vacuum studied in \cite{Karliner:1988hd}.}
All this is very elementary and its seems hard to believe that anything interesting could come out of it.
However, as we have remarked above, a local observer only has access to his own causal diamond, and thus cannot measure all of the Fourier modes of the scalar field.
The Penrose diagram of de Sitter is of course simply a rectangle, and the closer we are to future space-like infinity (the top of the diagram) the more causally disconnected regions there are on a constant $\tau$ (which implies also constant conformal time) slice. Let's consider such a slice with $\mathcal{O}(1/\epsilon)$ regions of typical size $\epsilon$. Our future observer that looks back long after inflation has ended can in principle measure the average state (expectation value and momentum) inside all of these boxes\footnote{In the case of a strict de Sitter universe inflation of course never ends, but we could imagine an unphysical, omniscient being that has access to the expectation values of the field inside all the causally disconnected regions on a space-like hypersurface. This latter point of view is of course not just unphysical in that no such observer exists, but it also opens up a Pandora's box of dicey questions concerning the interpretation of quantum mechanics in the context of cosmology. We will not discuss such issues here, so our discussion of de Sitter should be understood merely as an idealization approximating the post-inflation observer's point of view.}, and may choose to introduce corresponding coarse grained variables defined by convoluting the field profile with a suitable window function.
Let us choose for this purpose a Lorentzian window (or box) function centered at $\theta_i$
\[ \label{Box}
\mathcal{B}_i(\theta - \theta_i) = \sqrt{\epsilon^3 \over 4 \pi}\, {1 \over (\theta - \theta_i)^2 + (\epsilon/2)^2} \ ,
\]
since its Fourier transform is conveniently given by an exponential, and thus the average value of the field inside this box is given by the convolution
\[
\int_{-\infty}^{\infty} \mathrm{d}\theta\, \mathcal{B}_i(\theta - \theta_i)\, \phi(\theta) = \sqrt{ \pi\epsilon} \sum_{n=-\infty}^{\infty} e^{i n \theta_i} e^{-n \epsilon/2}\, \phi_n \ .
\]
Here we have taken the range of integration as infinite, which is an approximation that becomes exact in the limit of vanishing $\epsilon$ which we will be interested in below (or alternatively we can think of the integral as simply wrapping around the circle).
Note that while the vacuum (\ref{EucVac}) is a pure state in the full Hilbert space, this is no longer true in the reduced state space resulting from coarse-graining.
What we are going to be concerned with are the statistical properties of these coarse-grained variables, and it is their overlap distribution we would like to compute. First, however, we will have to define this distribution and introduce some techniques for calculating it.
\subsection{Review of Overlap Distributions}
Let us briefly recall the basic concept of an overlap distribution and some useful definitions, as explained in \cite{MPV,CG,DG, Parisi:1983dx} and the excellent recent lecture notes \cite{Denef:2011ee}.
In complex materials a thermodynamic state described by a density matrix $\rho$ may at low temperatures break ergodicity (i.e.~not all accessible microstates or phase space cells will be equally likely). This can occur as a trivial consequence of symmetry breaking (as in a ferromagnet), but in more non-trivial systems such as spin glasses ergodicity breaking takes place irrespective of any obvious associated symmetry. We should then be able to decompose the density matrix into different ergodic components (sometimes referred to as thermodynamic pure states)
$\rho = \sum_\alpha w_\alpha \rho_\alpha$.
Different ergodic components in phase space are separated by very high free energy barriers, and thus no longer communicate with each other at low temperatures.
The overlap $q_{\alpha\beta}$ between two such components is then defined by averaging over all degrees of freedom the product of their expectation values in states $\alpha$ and $\beta$.
E.g.~for a system of $N$ spins $s_i$, we would write $q_{\alpha\beta} \equiv {1 \over N} \sum_i \langle s_i \rangle_\alpha \langle s_i \rangle_\beta$.
Usually it is not possible to explicitly construct the decomposition into thermodynamic pure states, but a quantity of interest that may be computable is the overlap distribution
\[ \label{OverlapDef}
\mathcal{P}(q) \equiv \sum_{\alpha,\beta} w_\alpha w_\beta\, \delta(q - q_{\alpha \beta}) = \left\langle \delta\left( q - {1 \over N} \sum_i s_i^1 s_i^2 \right) \right\rangle_{2\ \mathrm{replicas}} \ .
\]
The last equality is a rather non-trivial step, that forms the centerpiece of the replica formalism. It re-expresses the sum over ergodic components (which we cannot perform explicitly) as an expectation value of an overlap between two copies of the system. We will not reproduce the proof of this statement here, but instead refer the reader to the literature for justification \cite{MPV, Denef:2011ee}.\footnote{Note that there is nothing unphysical about working with two copies of the same system, and that this should not be confused with the so-called replica trick for computing quenched averages of the free energy, which employs an arbitrary number of replicas and then analytically continues that number to zero.}
As we keep lowering the temperature, the breaking up of the phase space may continue in several steps, leading naturally to a tree structure in the space of states. This is best probed by computing the triple overlap distribution
\[
\mathcal{P}(q_1,q_2,q_3) \equiv \sum_{\alpha,\beta,\gamma} w_\alpha w_\beta w_\gamma \, \delta(q_1 - q_{\beta\gamma})\, \delta(q_2 - q_{\gamma \alpha}) \, \delta(q_3 - q_{\alpha \beta}) \ ,
\]
which can be similarly expressed in terms of a three replica expectation value. If a hierarchical tree structure is present in the space of states it will be reflected in this quantity exhibiting ultrametricity \cite{MV}, namely the property that out of any triple of overlaps $(q_1,q_2,q_3)$ the largest two will be equal, as is the case for pairwise path-distances between any three leaves of a tree. This means that the distribution will have support only on a codimension one subspace where the three $q_a$ form equilateral or isosceles triangles.
\subsection{Introducing Wigner Densities}
Now let's consider more generally $N$ degrees of freedom living in a Hilbert space $\mathcal{H}$, and denote by $\mathcal{H}_i$ the Hilbert space associated with one of them. We can find a reduced density matrix $\rho^i$ by tracing over all other degrees of freedom, and use this to define a new kind of overlap\footnote{We use the upper case symbol $Q_{\alpha\beta}$ here to stress the analogy with $q_{\alpha\beta}$ in the spin example, but note that the range in which this overlap takes values is shifted and rescaled compared to the definition above.}, first introduced in \cite{Denef:2011ee},
between two states labelled by $\alpha$ and $\beta$ as
$Q_{\alpha\beta} \equiv {1 \over N} \sum_i \mathrm{Tr}_{\mathcal{H}_i}\, \rho^i_\alpha \rho^i_\beta$.
We can recast this overlap in a very suggestive form by working with Wigner distributions instead of density matrices. Recall that the Wigner distribution in phase space is defined as a Fourier transform of certain matrix elements of the density operator. In the case at hand, for a scalar field $\phi$ with momentum $p$, in the Euclidean vacuum state $\rho = |\Psi\rangle \langle \Psi |$, we find
\[
W(\phi, p) \equiv \int \mathrm{d} s\, e^{i s.p}\, \langle \phi + s/2 | \rho | \phi - s/2 \rangle = \prod_{n=1}^\infty 4\, e^{-2n |\phi_n|^2 - |p_n|^2 /(2 n)} \ ,
\]
where $s$ and $p$ satisfy the same reality condition as $\phi$, and the integral measure is $\mathrm{d} s = \prod_{n=1}^\infty \mathrm{d} s_n$, with the $s_n$ integrated over the whole complex plane. Here we have set Planck's constant to unity and used the shorthand $s.p = {1 \over 2}\sum_{n = -\infty}^\infty s_n\, p_n$, again with the zero-mode removed.
We can also write down the Wigner densities $W^i$ corresponding to the reduced density matrices for the average expectation value $X_i$ and conjugate momentum $P_i$ of the field inside a certain region $\mathcal{B}_i$ \cite{Unpub}
\[ \label{RedWig}
W^i = \int \mathrm{d} \phi\, \mathrm{d} p \,\delta\left(X_i - \int_{-\infty}^{\infty} \mathrm{d}\theta\, \mathcal{B}_i(\theta - \theta_i)\, \phi(\theta)\right)\, \delta\left(P_i - \int_{-\infty}^{\infty} \mathrm{d}\theta\, \mathcal{B}_i(\theta - \theta_i)\, p(\theta)\right) \, W(\phi, p) \ , \quad
\]
where the integration measure is $(2 \pi)^{-2} \prod_{n=1}^\infty \mathrm{d} \phi_n \, \mathrm{d} p_n$ and the range again extends over the whole complex plane.\footnote{The coarse-graining involved in the definition of the reduced Wigner densities reflects the thermal nature of the cosmological horizon of a localized observer; in particular, the associated density matrices no longer correspond to pure quantum states such as the Euclidean vacuum state (\ref{EucVac}) we stated with.}
Using these reduced Wigner distributions on the phase space for one particular degree of freedom, we can recast the overlap in the following form \cite{Denef:2011ee}
\[
Q_{\alpha\beta} = {1 \over N} \sum_i \int \mathrm{d} X_i \, \mathrm{d} P_i \, W^i_\alpha(X_i, P_i) \, W^i_\beta(X_i, P_i) \ ,
\]
except that in the calculations below we will not actually work with $N$ fixed boxes in positions $\theta_i$ and average over them, but instead replace this sum by an integral $(2 \pi)^{-1} \int_0^{2\pi} \mathrm{d} \theta_i$.
This expression for the overlap can now be used to rewrite the overlap distribution as a two replica expectation value, in complete analogy with the right hand side of (\ref{OverlapDef}). While an expectation value of a delta function may appear to be a somewhat singular representation, the important implication is that the moments of the distribution \cite{Unpub} are given by
\[ \label{MomDef}
\langle q^K \rangle = \left(\prod_{i=1}^K \int_0^{2 \pi} {d \theta_i \over 2 \pi} \int_{-\infty}^\infty \mathrm{d} X_i \, \mathrm{d} P_i \right) \left(W^{123 \ldots K}\right)^2 \ ,
\]
where $W^{123 \ldots K}$ is a multivariable Wigner density defined exactly as in (\ref{RedWig}) except with $K$ pairs of delta function insertions
that set the average expectation values of the field $\phi(\theta)$ inside the boxes at positions $\theta_i$ equal to $X_i$ and similarly equate the average conjugate momenta to $P_i$.
For our particular choice of window function (\ref{Box}) we can find the reduced Wigner distribution by using the integral representation of the $K$ pairs of $\delta$-functions with integration variables $\lambda_i$ for expectation values and $\tilde{\lambda}_i$ for momenta, which gives
\[ \label{Wigner}
W^{123\ldots K} &=& \int_{-\infty}^{\infty} \left(\prod_{k=1}^K \, \mathrm{d}\lambda_k\, \mathrm{d}\tilde{\lambda}_k \, \ e^{2\pi i (\lambda_k X_k + \tilde{\lambda}_k P_k)} \right) \\
&&\times \exp \left( - 4 \pi^3 \epsilon \sum_{n=1}^{\infty} e^{- n \epsilon} \sum_{i,j = 1}^{K} \left({\lambda_i \lambda_j \over 2 n} + 2 n \tilde{\lambda}_i \tilde{\lambda}_j \right) \cos n (\theta_i - \theta_j) \right) \ . \nonumber
\]
Provided we can compute from (\ref{MomDef}) the moments for all integers $K$, we can find the characteristic function, and hence reconstruct the full overlap distribution $\mathcal{P}(q)$.
If the system has a non-trivial phase structure, which is related to failure of cluster decomposition as discussed in \cite{Denef:2011ee}, we should find a non-trivial overlap distribution. In this sense this quantity can be understood as an order parameter for ergodicity breaking.
\section{A First Stab at the Overlap Distribution} \label{Low}
Having explained the basic setup of the computation, let us now tackle the calculation of the first few moments of the overlap distribution.
From the analogy with complex materials in the previous section, it should be apparent that the box size $\epsilon$ plays a role analogous to temperature, and therefore we will be interested in the small $\epsilon$ (i.e.~late conformal time) expansion in which there are many causally disconnected regions in our model universe.
\subsection{First Moment}
In order to find the first moment of the overlap distribution, we merely have to simplify the expression (\ref{Wigner}) for the reduced Wigner density with one index. There are no angular integrals to be performed (since by rotational symmetry the moments only depend on differences of angles), and the integrals over $X_1$ and $P_1$ simply lead to delta functions that set to zero the sums of the $\lambda_i$ and $\tilde{\lambda}_i$ parameters, respectively, for the two copies of the Wigner density.
Evaluating the sums $\sum_{n=1}^{\infty} 2 n \, e^{- n \epsilon} = (\cosh \epsilon -1)^{-1}$ and $\sum_{n=1}^{\infty} e^{- n \epsilon}/(2n) = -{1\over 2} \log(1-e^{-\epsilon})$ we find that the first moment is given by
\[
\langle q \rangle &=& \int \mathrm{d} X_1 \mathrm{d} P_1 \, (W^1)^2 = \int_{-\infty}^{\infty} \, \mathrm{d}\lambda_1 \mathrm{d}\tilde{\lambda}_1 \exp \left( - 8 \pi^3 \epsilon \left( -{\lambda_1^2 \over 2} \log(1-e^{-\epsilon}) + { \tilde{\lambda}_1^2 \over \cosh \epsilon -1 } \right) \right) \nonumber \\
&=& { \sinh {\epsilon \over 2} \over 4 \pi^2 \epsilon \sqrt{ -\log(1-e^{-\epsilon})}} \ \stackrel{\epsilon \to 0}{ \longrightarrow } \ {1 \over 8 \pi^2 \sqrt{-\log \epsilon}} + \mathcal{O} (\epsilon(-\log\epsilon)^{-3/2}) \ .
\]
Note that the leading term is proportional to $(-\log\epsilon)^{-1/2}$, which tells us that the average overlap goes to zero (albeit rather slowly) with conformal time. The precise coefficient of $\langle q \rangle$ depends our choice of window function (i.e.~on how many boxes of a given shape and width $\epsilon$ can effectively fit into our one-dimensional universe), and thus is not particularly physical, though it will be crucial when comparing to higher moments of the distribution.
\subsection{Second Moment}
To compute higher moments we have to perform angular integrals over the (differences in) box positions.
For $K=2$ there is only one independent angle (since one can always be fixed using the rotational symmetry) and the integral can be evaluated exactly.
Using the sums
\[
\label{SumP}
\sum_{n=1}^{\infty} 2 n \, e^{- n \epsilon} \cos(n \Delta\theta) &=& {\cosh \epsilon \cos \Delta\theta -1 \over \cosh \epsilon - \cos \Delta\theta} \ , \\
\label{SumX}
\sum_{n=1}^{\infty} {e^{- n \epsilon} \over 2n} \cos(n \Delta\theta) &=& -{1\over 4} \log(1-2 e^{-\epsilon} \cos \Delta\theta + e^{-2\epsilon}) \ ,
\]
we find that the second moment is given by
\[
\int \mathrm{d} X_1 \mathrm{d} P_1 \mathrm{d} X_2 \mathrm{d} P_2 \, (W^{12})^2 = {1 \over 64 \pi^4 \epsilon^2} { 1 \over \sqrt{\det M}}{ 1 \over \sqrt{\det \tilde{M}}} \ ,
\]
where the symmetric matrix $\tilde{M}$ has entries $\tilde{M}_{ij}$ given by equation (\ref{SumP}) with $\Delta\theta = \theta_i - \theta_j$, and similarly the symmetric matrix $M$ has matrix elements given by equation (\ref{SumX}).\footnote{We should remark at this point that in principle any reasonable (i.e.~continuous and square integrable) window function could have done do the job instead of (\ref{Box}), and suitably regularized the computation of the matrix elements of $M$ and $\tilde{M}$. However, the Lorentzian is particularly convenient, since it allows us to explicitly perform the resulting sums (\ref{SumP}) and (\ref{SumX}) which makes the following computations much more manageable. We have tried out different box functions, and found that the crucial long-range behavior of (\ref{SumX}) was identical, but accompanied by various short range terms modifying this for small $\Delta\theta$. If we grant that the long range behavior is universal, the detailed choice of box function should not matter for subsequent calculations of the shape of the (appropriately rescaled) distribution in the small $\epsilon$ limit.}
Due to the smoothing provided by the window function, all matrix elements are finite, and divergences can only occur when one of the matrices develops a zero eigenvalue (such that the determinant vanishes) which happens when the two boxes coincide. This singularity is not integrable, thus we have to be careful to exclude from the integral a region in which the angular separation becomes less than $\mathcal{O}(\epsilon)$.\footnote{A different method to treat coincident boxes, which may be conceptually clearer, is as follows. Instead of defining the the reduced Wigner distribution with an insertion of $\delta(X_i - \int \mathrm{d} \theta \, \mathcal{B}_i \phi(\theta))$, we can broaden the delta function to a Gaussian of width $\delta$, i.e. instead of demanding that the averaged field be exactly equal to a specified value, we only ask for approximate (smeared) agreement. It is easy see that this ensures that matrix $M$ never degenerates, i.e. has non-vanishing determinant even for coincident boxes. The same is true for the matrix $\tilde{M}$ with the corresponding broadening of the momentum space delta function. We should then compute the moments of the overlap distribution as functions of $\epsilon$ and $\delta$ and in the end take the double limit of both small parameters going to zero. Unfortunately this makes the calculation somewhat unwieldy in practice.}
It is not hard to see that the momentum determinant can be expanded in a power series in $\epsilon$
\[
{ 1 \over \sqrt{\det \tilde{M}}} = {\epsilon^2 \over 2} + \mathcal{O}(\epsilon^4) \ .
\]
Furthermore, even though subleading terms in this expression diverge for vanishing angular separation $\Delta\theta =0$, after performing the angular integral over the interval $(\epsilon, 2\pi - \epsilon)$, say, the contribution of the momentum determinant remains $\epsilon^2/2$ to leading order.
Thus we have to focus on the position determinant, and we will see below that this behavior continues for the computation of higher moments: the momentum determinant contributes only a constant, since the momentum correlators are short-ranged.
It is convenient to rescale the parameters $\lambda_i$ by $(-4\pi^3 \epsilon \log(1-e^{-\epsilon}))^{-1/2}$ so that the diagonal terms become $\epsilon$ independent Gaussians, and
\[ \label{qsquared}
\langle q^2 \rangle = {1 \over 8 \pi^2 \epsilon} {\epsilon^2 \over 2} \int {\mathrm{d} \Delta\theta \over 2 \pi} \int_{-\infty}^{\infty} \, {\mathrm{d}\lambda_1 \mathrm{d} \lambda_2 \, e^{-\lambda_1^2 -\lambda_2^2 } \over - 4 \pi^3 \epsilon \log(1-e^{-\epsilon}) } \left( 1 - 2 e^{-\epsilon} \cos \Delta\theta + e^{-2\epsilon}\right)^{\lambda_1 \lambda_2 \over -\log(1-e^{-\epsilon})} \ .
\]
We are now supposed to integrate over the parameters $\lambda_i$, and in the end over the angle $\Delta\theta$, being careful to exclude a small region around $\Delta\theta=0$. This is not very practical however, and we will instead switch the order of integration.
Performing the angular integral first, over the full range $(0,2\pi)$, which in this case can be done explicitly, leads to a hypergeometric function depending on $\epsilon$ and the combination $c \equiv {-\lambda_1 \lambda_2 / \log(1-e^{-\epsilon})} $ appearing in the exponent above. Of course, if we then try to integrate over the parameters $\lambda_i$ we again encounter the singularity originating from coincident boxes. For very negative values of the exponent $c$ the hypergeometric function grows so fast that the Gaussian envelope in (\ref{qsquared}) fails to render the integral finite.
It would appear that we have not gained anything, and still have to introduce a regulator to obtain a well-defined answer. However, at this point we recall that we are really interested in the thermodynamic limit, which in this case means the limit of small $\epsilon$. At late cosmological times, the (causally disconnected) boxes become very small, and two randomly chosen ones are very unlikely to be coincident.
Taking the small $\epsilon$ limit after carrying out the angular integral in (\ref{qsquared}) allows us to separate the divergent part of the answer from the $\mathcal{O}(\epsilon^0 (\log \epsilon)^n)$ terms that we are interested in.\footnote
{
More precisely, if we consider the resulting hypergeometric function for small $\epsilon$, but keeping $c$ constant, the expansion we find is not uniform. It is the sum of a power series starting at $\mathcal{O}(\epsilon^0)$ and another power series that contains a factor $\epsilon^{2c}$ and starts at $\mathcal{O}(\epsilon^{2c+1})$. The latter contains the singularity at large negative $c$, and we drop it, while the former describes the behavior at small $c$ (it dominates for $c > -1/2$), and thus we keep its leading term. Remembering that $c$ goes to $-\lambda_1\lambda_2/\log \epsilon$ it is clear that the physically interesting region is that of small $c$.
Put differently, the terms that lead to divergences in the $\lambda_i$ integral are $\mathcal{O}(\epsilon^1)$ with a prefactor $e^{-2 \lambda_1 \lambda_2}$, while the terms we are interested are present at zeroth order in the $\epsilon$ expansion (with correction in powers of $\log \epsilon$), but don't contain the offending exponential factor.
}
Note that of course one cannot simply set $\epsilon$ to zero from the start. It is crucial to extract the dependence on $\log \epsilon$ first, which is most easily done by rescaling the $\lambda_i$ appropriately. In fact, going back to (\ref{qsquared}) one can check that performing the integral for finite $\epsilon$ and then taking the thermodynamic limit is equivalent to letting $\epsilon$ go to zero in the integrand, such that it appears only inside inside the logarithms, and then performing the integral.
This leads to an expansion of the second moment in inverse powers of $\log\epsilon$
\[
\langle q^2 \rangle = { 1 \over -64 \pi^4 \log\epsilon} \left( 1+ {\pi^2 \over 24 \log^2 \epsilon} + {19 \pi^4 \over 640 \log^4 \epsilon} +{1375 \pi^6 + 151200 \zeta^2(3) \over 21504 \log^6 \epsilon} + \mathcal{O}(\log^{-8} \epsilon )\right) \ . \qquad
\]
Higher order terms are easily computed by expanding further. Note that this expansion satisfies a transcendentality principle; each inverse power of $\log\epsilon$ comes with a power of $\pi$ (or corresponding $\zeta$ value). Also, the leading term is precisely equal to $\langle q \rangle^2$.
In summary, our prescription for computing the overlap, which we will also apply to higher moments below, is to first rescale the auxiliary parameters $\lambda_i$ and take the thermodynamic limit, throwing away terms proportional to powers of $\epsilon$ (which may have led to divergences), but being careful to retain the dependence on $\log \epsilon$. After that, we perform the angular integral(s) over the full interval from $0$ to $2\pi$, and in the end integrate over the (rescaled) $\lambda_i$.
\subsection{Third Moment}
Computing higher moments of the overlap distribution requires us to perform multiple angular integrals, while being careful to treat coincident boxes properly. This can again be formulated as finding the determinants of $M$ and $\tilde{M}$, which for the $K$th moment are $K$ by $K$ matrices, with the off-diagonal matrix elements $M_{ij}$ and $\tilde{M}_{ij}$ depending on the angular separation $\theta_i - \theta_j$, again given by (\ref{SumX}) and (\ref{SumP}).
Hence for $K=3$ we have
\[
\int \prod_{i=1}^3 \mathrm{d} X_i \mathrm{d} P_i \, (W^{123})^2 = {1 \over (8 \pi^2 \epsilon)^3} { 1 \over \sqrt{\det M}}{ 1 \over \sqrt{\det \tilde{M}}} \ .
\]
As before one can show that the momentum determinant contributes only a constant factor, in this case ${ 2^{-3/2}\epsilon^3}$. Rescaling the $\lambda_i$ and taking the $\epsilon \to 0$ limit then leads to the expression
\[ \label{qcubed}
\langle q^3 \rangle &=& {1 \over (4 \pi)^3} \int_0^{2\pi} {\mathrm{d} \theta_1 \mathrm{d} \theta_2 \mathrm{d} \theta_3 \over (2 \pi)^3} \int_{-\infty}^{\infty} \, {\mathrm{d}\lambda_1 \mathrm{d} \lambda_2 \mathrm{d} \lambda_3 \, \over (- 4 \pi^3 \log \epsilon)^{3/2} } \ e^{-\lambda_1^2 -\lambda_2^2 -\lambda_3^2 } \nonumber \\
&&\times \left(4\sin^2 { \theta_1-\theta_2 \over 2}\right)^{-{\lambda_1 \lambda_2 \over \log \epsilon}} \left(4\sin^2 { \theta_1-\theta_3 \over 2}\right)^{-{\lambda_1 \lambda_3 \over \log \epsilon}} \left(4\sin^2 { \theta_2-\theta_3 \over 2}\right)^{-{\lambda_2 \lambda_3 \over \log \epsilon}} \ .
\]
While it is not obvious a priori how to compute this integral exactly for arbitrary exponents, we can try to use a procedure similar in spirit to the replica trick: find an appropriate function that agrees with this integral for integer exponents (in which case the integral can be performed explicitly) and then use this function to extrapolate to small values of the exponent. This is discussed in Appendix \ref{K3Int}.
Using the result (\ref{3AngInt}) from the appendix we find that the third moment is given by
\[
\langle q^3 \rangle = {1 \over (-64 \pi^4 \log\epsilon)^{3/2}} &&\left( 1+ {\pi^2 \over 8 \log^2 \epsilon} + {\zeta(3) \over 4 \log^3 \epsilon} + {67 \pi^4 \over 640 \log^4 \epsilon} \right. \nonumber \\
&&\left. +{9 \pi^2 \zeta(3) + 108 \zeta(5) \over 32 \log^5 \epsilon}+ \mathcal{O}(\log^{-6} \epsilon )\right) \ .
\]
Again this obeys the transcendentality principle, and one can easily compute further terms if one so desires.
\subsection{The Lowest Central Moments}
We have already noted that to leading order $\langle q^2 \rangle = \langle q \rangle^2$ and thus the second central moment (variance) is actually of order $\log^{-3} \epsilon$:
\[
\langle q^2 \rangle_c = {1 \over -64 \pi^4 \log\epsilon} \left( {\pi^2 \over 24 \log^2 \epsilon} + {19 \pi^4 \over 640 \log^4 \epsilon} + \mathcal{O}(\log^{-6} \epsilon ) \right ) \ .
\]
More surprisingly, computing the third central moment we find that the first two leading terms cancel out
\[
\langle q^3 \rangle_c = \langle q^3 \rangle - 3\langle q^2 \rangle \langle q \rangle +2 \langle q \rangle^3 = {1 \over (-64 \pi^4 \log\epsilon)^{3/2}} \left( {\zeta(3) \over 4 \log^3 \epsilon} + {\pi^4 \over 64 \log^4 \epsilon} + \mathcal{O}(\log^{-5} \epsilon )\right) \ . \qquad
\]
This raises the question of whether the $K$th central moments $\langle q^K \rangle_c$ are all of $\mathcal{O}(\log^{-3K/2} \epsilon)$.
If this is the case there might be a closed form expression for the leading overlap distribution (i.e. capturing the leading term of each central moment) as a function of $\hat{q} \sim (\log^{3/2} \epsilon) (q - \langle q \rangle) $, without explicit $\epsilon$ dependence.
In fact, the above cancellations suggests that perhaps one should set up the calculation directly for central moments, rather than computing ordinary moments and subtracting. Certainly, if we were to use numerical methods this would be absolutely necessary, and
we will see below that this also leads to substantial simplifications in the exact analytic computation.
\subsection{Fourth and Higher Moments}
The computation of higher moments is formally very similar to what we have discussed above.
The crucial integral that needs to be evaluated to find the $K$th moment of the overlap distribution is given by\footnote{Essentially this amounts to studying a one-dimensional model of particles on a circle with a pairwise interaction potential given by the logarithm of the chordal distance, and arbitrary real-valued charges $\lambda_i$.}
\[ \label{HigherInt}
\int_0^{2\pi} \prod_{k=0}^K {\mathrm{d}\theta_k \over 2 \pi} \prod_{j>i \ge 1}^K \left(4 \sin^2 { \theta_i-\theta_j \over 2}\right)^{-\lambda_i \lambda_j / \log\epsilon} \ .
\]
Knowledge of this integral for all positive integers $K$, even if only as an expansion for small $\epsilon$, would allow us to reconstruct the complete overlap distribution.
However, already for $K=4$ this is rather non-trivial. Appendix \ref{K4Int} describes the treatment analogous to the one that worked for $K=3$, while Appendix \ref{Modest} discusses what we get if we expand in $\epsilon$ first. It is clear from these discussions that it would be difficult to directly compute arbitrary higher moments in this fashion, and we instead have to think of a smarter method to achieve this.
\section{Central Moments and Graphs} \label{Graphs}
From the above we know that the overlap distribution is approximately a delta-function centered on $\langle q \rangle = (-64 \pi^4 \log\epsilon)^{-1/2}$, and that we should really compute central moments to learn more about its shape.
\subsection{Expansion in Terms of Complete Graphs}
The central moments are given by
\[
\langle q^K \rangle_c &\equiv& \langle (q-\langle q \rangle)^K \rangle = \sum_{n=0}^K {K \choose n} \langle q^n \rangle (-\langle q \rangle)^{K-n} \\
&=& \langle q^K \rangle - K \langle q^{K-1} \rangle \langle q \rangle + {1\over 2} K (K-1) \langle q^{K-2} \rangle \langle q \rangle^2 - \ldots + (-1)^{K-1} (K-1) \langle q \rangle^K \ . \nonumber
\]
Every term in this expansion, when written out explicitly in terms of Wigner functionals as in the previous section, will boil down to a certain angular integral of the type discussed above, which we can represent symbolically by a graph. If draw a vertex labelled by $i$ for the angular integral $\int \mathrm{d} \theta_i / (2 \pi) $ and an (undirected) edge between vertices $i$ and $j$ for each factor of $(4 \sin^2 { \theta_i-\theta_j \over 2})^{-\lambda_i \lambda_j / \log\epsilon}$, then $\langle q^K \rangle$ corresponds to the complete graph with $K$ vertices (i.e. the graph in which each vertex is connected to every other one by exactly one undirected edge). We will denote this graph by $\mathcal{C}_K$.
The next term in the expansion above has a factor of $\langle q \rangle$, which corresponds to an integral over an angle that nothing depends on, which is represented by a vertex that has no edges attached to it. This gives simply a pure number, which after factoring out a suitable overall coefficient is unity. The factor $\langle q^{K-1} \rangle$ on the other hand again corresponds to a complete graph, but with only $K-1$ vertices. There are $K$ such terms, and we can think of these as arising from the $K$ complete graphs $\mathcal{C}_{K-1}$ that are subgraphs of $\mathcal{C}_K$.
Similarly, all subsequent terms in the above expansion can be associated to complete graphs of fewer vertices, and the combinatorics is such that there is precisely one term of coefficient one or minus one for every complete graph that is a subgraph of $\mathcal{C}_K$. Finally, the last term is just a number, $(-1)^{K-1} (K-1)$, indicating graphs with no edges, which we can think of as arising from $K$ complete graphs with one vertex (singletons) and one graph with no vertices (the null graph).
Why is this relevant? Naively the $K$th moment will be of $\mathcal{O}(\log^{-K/2} \epsilon)$, but we have seen above that due to certain cancellations at the lowest central moments are actually parametrically smaller than that. The graphical representation can help us rewrite the computation in a way the makes these cancellations manifest for all $K$, and allows us to easily identify the leading pieces of the answer in the small $\epsilon$ limit.
\subsection{Expansion with Parametrically Small Edge Factors}\label{EdgeExp}
We will illustrate the idea with the simple example of $K=3$. If we introduce the shorthand $(ij) \equiv (4 \sin^2 { \theta_i-\theta_j \over 2})^{-\lambda_i \lambda_j / \log\epsilon}$ for the angular factor associated to the edge between vertices $i$ and $j$, and declare it as understood that all angles will be integrated over with measure $\int_0^{2 \pi} \mathrm{d} \theta_i / (2 \pi) $, and all parameters $\lambda_i$ with measure $\int_{-\infty}^{\infty} \mathrm{d} \lambda_i \exp(-\lambda_i^2) / (\sqrt{\pi}) $, we have
\[
(-64 \pi^4 \log\epsilon)^{3/2} \langle q^3 \rangle_c &=& (12)(13)(23) - (12) - (13) - (23) + 2 \\
&=& [(12)-1][(13)-1][(23)-1] \nonumber \\
&+& [(12)-1][(13)-1] + [(12)-1][(23)-1] + [(13)-1][(23)-1] \ . \nonumber
\]
Here we have rewritten the polynomial in $(ij)$ in terms of shifted variables, which we will denote by $[ij] \equiv (ij)-1$. While the $(ij)$ are of $\mathcal{O}(1)$, the factors of $[ij]$ are of $\mathcal{O}(\log^{-1}\epsilon)$, which makes it obvious that the right hand side is actually of $\mathcal{O}(\log^{-3}\epsilon)$ and that the leading contribution comes entirely from the term cubic in the rectangular brackets.
To see this, we merely have to consider the small $\epsilon$ expansion of the $[ij]$. In the first (cubic) term we can expand each factor to first order in $\log^{-1}\epsilon$ and the leading term will be proportional to $\lambda_1^2 \lambda_2^2 \lambda_3^2\,$, which makes a non-vanishing contribution under the $\lambda_i$ integrals. In the remaining (quadratic) terms however, we have to the expand each factor of $[ij]$ to second order in $\log^{-1}\epsilon$, since the first order terms are odd in some $\lambda_i$ and thus vanish. Hence the contribution of these terms is subleading and of order $\log^{-4}\epsilon$.
It pays to again think of this in terms of graphs, except that now the edge factors are equal to $[ij]$. The first (cubic) term corresponds to a triangle (namely $\mathcal{C}_3$), and its leading contribution comes from the shortest loop\footnote{In graph theory this would be called a closed walk, since the word loop is reserved for an edge that connects a vertex to itself. There are no such loops in our graphs, so hopefully this abuse of terminology will not be confusing.} we can draw on a triangle (namely going around the triangle once) which corresponds to expanding each factor to first order. The quadratic terms on the other hand correspond to graphs with only two edges, and they contribute only once we expand every edge to second order, which again corresponds to the shortest loop we can draw on such a graph, namely going along the two edges and then back again. However this loop is of length four, and thus subleading compared to the triangle, which has a loop of length three.
This will be the general theme: we will write the result for $\langle q^K \rangle_c $ as a sum over terms labelled by certain graphs, and the contribution of each term will be suppressed by as many powers of $\log^{-1}\epsilon$ as the length shortest link (set of loops) we can draw on this graph.
As a slightly more non-trivial example, consider $K=4$. Again the expansion in terms of the $(ij)$ contains the complete graph $\mathcal{C}_4$ and all its complete subgraphs $\mathcal{C}_{4-n}$ with coefficient $(-1)^{4-n}$, accounting for multiplicities arising from the fact that we consider the vertices as labelled, i.e. distinguishable:
\[
&& (-64 \pi^4 \log\epsilon)^{2} \langle q^4 \rangle_c \\
&& \ = (12)(13)(14)(23)(24)(34) - (12)(13)(23) - (12)(14)(24) \nonumber \\
&& \ - (13)(14)(34) - (23)(24)(34) + (12) + (13) + (14) + (23) + (24) + (34) - 3 \nonumber \\
&& \ = [12][13][14][23][24][34] \nonumber \\
&& \ + [13][14][23][24][34] + (\mathrm{five\ similar\ terms\ with\ five\ factors}) \nonumber \\
&& \ + [14][23][24][34] + (\mathrm{fourteen\ similar\ terms\ with\ four\ factors}) \nonumber \\
&& \ + [14][24][34] + (\mathrm{fifteen\ similar\ terms\ with\ three\ factors\ and\ without\ closed\ loops}) \nonumber \\
&& \ + [12][34] + [13][24] + [14][23] \nonumber \ .
\]
The $[ij]$ expansion on the right hand side has a more interesting structure. It contains $\mathcal{C}_{4}$ which has six edges, all of its subgraphs with five edges and also every subgraph with four edges. Out of the 20 possible subgraphs with three edges, there are 16 present, and the four that are missing are precisely the complete graphs $\mathcal{C}_{3}$, i.e. triangles. Similarly, all subgraphs of these four $\mathcal{C}_{3}$ are absent and thus there no graphs with one or zero edges. The only graphs with two edges are the three that are not subgraphs of any of the $\mathcal{C}_{3}$. All those that appear do so with unit coefficient.
In other words, the right hand side contains exactly one term for every subgraph of $\mathcal{C}_{4}$ which has at least one edge attached to every vertex (of $\mathcal{C}_{4}$), i.e. has no isolated vertices.
Which terms give the leading contribution in this case? It is clear that the shortest link we can draw on a graph with $K$ vertices none of which are isolated, in such a way that every edge is traversed at least once, is of length $L_{\mathrm{tot}}=K$. In particular, if the expansion above contained $\mathcal{C}_{3}$, there would be loops of length $L=3$ covering that graph, but since this is not the case we have to look for loops of length $L=4$. Among the 15 graphs with four edges there are three containing a loop of length four. Furthermore, the last line contains three graphs which consists of only two edges. Each of those edges is completely disconnected from the rest of the graph, and we will refer to these as dimers. On a dimer we can draw a loop of length $L=2$ by going back and forth along the same edge, and we will call such loops trivial. Thus the three graphs in question also support links of length $L_{\mathrm{tot}}=4$, namely two mutually disconnected trivial loops each. Altogether only these six graphs, shown in Figure~1, will contribute at the leading order $\mathcal{O}(\log^{-4}\epsilon)$, which makes it much simpler to extract the dominant terms in the fourth central moment.
\begin{figure}
\begin{center}
\vspace{-1cm}
\includegraphics[width = 0.8 \textwidth]{6LK4G}
\vspace{-0.5cm}
\caption{\small The six graphs that contribute to the fourth central moment at leading order. For illustrative purposes we have chosen to arrange the four (distinguishable) vertices equidistantly on a circle, which can be thought of a constant time slice of $dS_2$.}
\vspace{-0.5cm}
\end{center}
\end{figure}
For general $K$ we can easily write down the $(ij)$ expansion. It turns out that the overall factor we need to extract is equal to $(-64 \pi^4 \log\epsilon)^{-K/2}$. Thus we can write $\langle q^K \rangle_c$ as a sum over terms labelled by complete subgraphs of $\mathcal{C}_{K}$, including singleton and null graphs.
\[ \label{CombRound}
(-64 \pi^4 \log\epsilon)^{K/2} \langle q^K \rangle_c
= \sum_{\mathrm{complete\, graphs\, with}\, n\, \mathrm{vertices}\, \subset \,\mathcal{C}_{K} } (-1)^{K-n} \prod_{j>i\ge1}^n (ij) \ . \qquad
\]
Our proposal is that when expanded in terms of the $[ij]$ this is equal to the much more useful expression
\[ \label{CombLemma}
(-64 \pi^4 \log\epsilon)^{K/2} \langle q^K \rangle_c
= \sum_{\mathrm{graphs}\, \mathcal{G} \mathrm{with}\, K\, \mathrm{non-isolated\, vertices}\, \subset \,\mathcal{C}_{K} } \ \ \prod_{\mathrm{edges}\, \overline{ij}\, \in\, \mathcal{G}} [ij] \ . \qquad
\]
We have checked explicitly that this holds also for $K=5$. The equality of the two graphical expansions on the right hand sides of these equations should be a simple consequence of the relation $[ij] \equiv (ij)-1$ (independent of the precise nature of the edge factor).\footnote{An iterative proof of this lemma for general $K$ might proceed along the following lines:
First one considers just the term corresponding to the complete graph with $K$ vertices, i.e.~the product of all ${1 \over 2} K (K-1)$ round bracket edge factors $\prod_{\overline{ij}\, \in\, \mathcal{C}_{K} } (ij)$, and expands this in terms of rectangular brackets. It is not hard to see that this expansion contains a term $\prod_{\overline{ij}\, \in\, \mathcal{G}} [ij]$ with unit coefficient for every subgraph $\mathcal{G}$ of the completely graph $\mathcal{C}_K$, including the complete graph itself. This is because multiplying out products of the $(ij)-1$, in order to express everything in terms of round brackets, and summing over all subgraphs, only the desired term associated with the complete graph survives. All other terms come with vanishing coefficient, since for terms with $r$ edges removed from a total of ${1 \over 2} K (K-1)$ we have $\sum_{p=0}^r (-1)^p {{1 \over 2} K (K-1) \choose p}{{1 \over 2} K (K-1) -p \choose r-p} = 0$.
Now in the round bracket expansion (\ref{CombRound}) we are interested in, terms corresponding to the $K$ complete (sub)graphs with $K-1$ vertices are subtracted from this. We can use the same identity again for each of them to express this as a rectangular bracket expansion containing terms with unit coefficient for every subgraph of $\mathcal{C}_{K-1}$. Subtracting these corresponds to removing all terms that have exactly one isolated vertex (namely the one not contained in $\mathcal{C}_{K-1}$) from the rectangular expansion.
However, terms that have two isolated vertices have now been subtracted twice, so we compensate for this we add to the round bracket expansion the complete subgraphs $\mathcal{C}_{K-2}$ that do not contain these vertices, and so on until we get down to the singleton and null graphs. We end up with a round bracket expansion that is an alternating sum of complete (sub)graphs versus a rectangular bracket expansion that contains all subgraphs except those with isolated vertices.}
After all, it just an algebraic statement that two polynomials are equal, and the graphs simply serve as a convenient way of labeling the terms in the two expansions.
\subsection{The Contribution of Cycle Graphs}\label{CycleGr}
Why is the latter expression (\ref{CombLemma}) so much more useful? We have already seen that it makes it obvious that the $K$th central moment will in fact be $\mathcal{O}(\log^{-3K/2}\epsilon)$, and allows us to easily identify a small subset of terms that contribute to the leading order. This is because the only graphs that contribute at this order are those that can be covered by links of length $L_{\mathrm{tot}}=K$ (i.e.~on which we can draw a set of loops using every edge of the graph at least once such that in total the number of steps does not exceed the number of vertices). It is easy to convince oneself that the relevant graphs are always composed of a set of (mutually disconnected) cycle graphs (in which every vertex has degree two) and dimers (two vertices of degree one joined by a single edge).
Not only are these graphs easy to identify, but we can also calculate their contribution explicitly. For a cycle graph this is much simpler than for a large connected graph (which we would have to compute in order to find the $K$th moment directly), since every vertex is connected only to two instead of $K-1$ other vertices.
Using the Fourier expansion of the logarithm of $(ij)$ given in (\ref{FourierEdge}) we find
\[
&&\int_0^{2\pi} {\mathrm{d} \theta_i \over 2 \pi} \left[\left(4 \sin^2 {\theta_{i-1}-\theta_i \over 2}\right)^{-\lambda_{i-1}\lambda_i/\log{\epsilon}} - 1\right] \left[\left(4 \sin^2 {\theta_{i}-\theta_{i+1} \over 2}\right)^{-\lambda_{i}\lambda_{i+1}/\log{\epsilon}} - 1\right] \nonumber \\
&=& {\lambda_{i-1} \lambda_i^2 \lambda_{i+1} \over \log^2 \epsilon} \sum_{n=1}^\infty {2 \over n^2} \cos(n(\theta_{i-1} - \theta_{i+1})) + \mathcal{O}(\log^{-3}\epsilon) \ .
\]
Concatenating further edge factors with this expression is trivial, since the result is self-similar - it simply leads to further factors of $\lambda_{j}\lambda_{j+1} \log^{-1} \epsilon$ and increases the power of $n$ in the denominator that multiplies the cosine of the difference between the first and the last angle. Once we close the loop after $L$ steps, there will be precisely two powers of $\lambda_i$ for every vertex we passed through, and the sum of $n^{-L}$ will result in a zeta value $\zeta(L)$. Thus if we identify $\theta_L = \theta_0$ and $\lambda_L = \lambda_0$ we find for $L \ge 3$
\[ \label{ZetaValue}
\prod_{i=1}^L [i-1,i]
&=& \int_{-\infty}^{\infty} \left(\prod_{i=1}^L {\mathrm{d}\lambda_i \over \sqrt{\pi}}\, e^{-\lambda_i^2}\right) \int_0^{2\pi} \left(\prod_{j=1}^L {\mathrm{d}\theta_j \over 2\pi} \right)\, \prod_{k=1}^L \left[\left(4 \sin^2 {\theta_{k-1}-\theta_k \over 2}\right)^{-\lambda_{k-1}\lambda_k/\log{\epsilon}} - 1\right] \nonumber\\
&=& (2 \log \epsilon)^{-L}\, 2\, \zeta(L) + \mathcal{O}(\log^{-L-2} \epsilon)\ .
\]
For trivial loops on dimers ($L=2$) the same result holds, except that there is an extra factor of $1/2$ since in this case we use the same edge twice and therefore have to expand to second order.
Thus a loop of length $L$ contributes a factor proportional to $\zeta(L)$. For a graph consisting of a number of disconnected cycles and dimers we simply multiply the contribution from each component. Since the total length of all loops in the link has to add up to $K$, this nicely confirms that the leading term of the (rescaled) $K$th central moment will have transcendentality $K$.
\subsection{Counting Graphs}
We have identified the graphs that contribute to the leading term of the central moments, and calculated the individual contributions of such graphs. What remains to be done is to count how many graphs will contribute for a given $K$. We will do this in two steps. First we need to enumerate the different families of graphs that are relevant, and then count the number of graphs within each given family.
We distinguish families simply according to how many (mutually disconnected) loops of length $L$ are present in the graph. We can characterize this by a set of integers $m_L$ that give the multiplicities of loops of a given length. In particular, $m_2$ gives the number of dimers, and the $m_L$ for $L>2$ count the number of cycles of length $L$ contained in the graph. We have already argued above that the graphs relevant to the leading central moments must obey the constraint
\[
\sum_{L \ge 2}\, m_L L = K \ .
\]
In other words, the families of graphs relevant for the $K$th central moment are labelled by the integer partitions of $K$ with the additional constraint that each summand has to be at least two.\footnote{This is known as an intermediate function in number theory.} We will denote the space of such integer partitions by $\mathsf{P}(2 , K)$.
How many graphs are there in a given family? If there are $m_L$ cycles of a given length $L$ we can begin by choosing $m_L$ indistinguishable sets of $L$ out of $K$ vertices, which can be done in
\[
{1 \over m_L!} {K \choose L} {K-L \choose L} \ldots {K - (m_L-1)L \choose L} = {K! \over m_L!\, (L!)^{m_L}\, (K- m_L L)! }
\]
ways. E.g.~we can choose to start with $L=2$ and then pick $m_3$ indistinguishable sets of three vertices out of the remaining $K - 2 m_2$ vertices and so on, which results simply in a product of factors of the form given above with $K$ replaced by the number of vertices remaining after each step.
Once we have picked sets of vertices we have to count how many different ways there are to make a cycle out of $L$ vertices. It easy to see that this is $L!$ modulo discrete rotations and orientation reversal, i.e. $(L-1)!/2$.
Again, the $L=2$ dimer is a special case, since there is clearly exactly one way of connecting to vertices with one edge, i.e.~the formula for cycles must be amended by an extra factor of 2.
In summary, for a given integer partition $m_L \in \mathsf{P}(2 , K)$, which specifies the family, the number of different graphs is given by
\[ \label{GraphCount}
K!\, 2^{m_2} \prod_{L \ge 2} {1 \over m_L! (2L)^{m_L}} \ .
\]
\subsection{Explicit Expressions for Central Moments}
Putting everything together, the leading central moments that follow from (\ref{CombLemma}) by summing over all families of graphs labelled by integer partitions, making contributions (\ref{ZetaValue}) with multiplicities (\ref{GraphCount}), are given by
\[ \label{CentralMoments}
\langle q^K \rangle_c = (-64 \pi^4 \log\epsilon)^{-K/2} (2 \log \epsilon)^{-K} K! \, \sum_{m_L \in \mathsf{P}(2 , K)}\, \prod_{L \ge 2} {1 \over m_L! } \left({\zeta(L) \over L}\right)^{m_L} + \mathcal{O}(\log^{-3K/2-1}\epsilon) \ . \qquad
\]
Happily, the extra factors of $1/2$ from having to expand to second order for dimers precisely cancel the extra factors of 2 from counting graphs with dimers, so that the $L=2$ case appears no different than $L \ge 3$ in the end.
This expression agrees with the leading terms in $\langle q^2 \rangle_c$ and $\langle q^3 \rangle_c$ found above, and gives the following results for the subsequent central moments
\[
\langle q^4 \rangle_c &=& (-256 \pi^4 \log^3 \epsilon)^{-2}\ {3 \pi^4 \over 20} + \mathcal{O}(\log^{-7} \epsilon) \ , \\
\langle q^5 \rangle_c &=& (-256 \pi^4 \log^3 \epsilon)^{-5/2}\ \left({10 \pi^2 \over 3} \zeta(3) + 24 \zeta(5) \right) + \mathcal{O}(\log^{-17/2} \epsilon) \ , \\
\langle q^6 \rangle_c &=& (-256 \pi^4 \log^3 \epsilon)^{-3}\ \left({61 \pi^6 \over 168} + 40 \zeta(3)^2 \right) + \mathcal{O}(\log^{-10} \epsilon) \ ,
\]
and so on. Here we have made use of the fact that zeta functions of even argument are expressible as rational numbers times powers of $\pi$.
As remarked above the central moments have uniform transcendentality and are $\mathcal{O}(\log^{-3K/2} \epsilon)$. This motivates us to define a new variable
\[
\hat{q} \equiv (-256 \pi^4 \log^3 \epsilon)^{1/2} (q - \langle q \rangle) \ ,
\]
in terms of which the overlap distribution will be smooth in the small $\epsilon$ limit.
\subsection{Characteristic Function and Overlap Distribution}
Knowing all central moments (at least to leading order) we would now like to reconstruct the overlap distribution. This is best done via the characteristic function, which is the expectation value of a complex exponential with frequency $\omega$.
Our result (\ref{CentralMoments}) involves a sum over integer partitions, so it is helpful to know that the generating function of (unrestricted) partitions is given by $\prod_{L = 1}^\infty (1-x^L)^{-1}$. We can easily modify this for our restricted case in which each summand has to be at least two, by simply omitting the $L=1$ factor. However, this is still not exactly what we need, since we are interested in integer partitions that come with additional factors of $1/m_L!$ depending on the multiplicities with which the integers $L$ appear. Fortunately for us, this case is even easier: the generating function is simply a product of exponentials, and thus
\[ \label{character}
\left\langle e^{i \omega \hat{q}} \right\rangle &=& \sum_{K=0}^\infty \langle\, \hat{q}^K \rangle\, { (i \omega)^K \over K! } = \prod_{L \ge 2} \exp\left((i \omega)^L {\zeta(L) \over L}\right) \nonumber \\
&=& e^{-i \gamma \omega}\ \Gamma(1 - i \omega) \ ,
\]
where we have used $\sum_{L \ge 2} (i \omega)^L \zeta(L)/L = - i \gamma \omega + \log(\Gamma(1 - i \omega))$ and $\gamma$ is the Euler-Mascheroni constant.
Using the integral representation of the $\Gamma$-function we can then Fourier transform the characteristic function to obtain the original overlap distribution for the (shifted and rescaled) variable $\hat{q}$:
\[ \label{Gumbel}
\mathcal{P}(\hat{q}) &=& {1 \over 2 \pi} \int_{-\infty}^{\infty} \mathrm{d} \omega \, e^{-i \omega \hat{q}} \left\langle e^{i \omega \hat{q}} \right\rangle = \int_0^\infty \mathrm{d} t \, e^{-t}\, \delta(\hat{q} + \gamma + \log{t}) \nonumber \\
&=& \exp(-\hat{q}-\gamma-e^{-\hat{q}-\gamma}) \ .
\]
This is known as a Gumbel or log-Weibull distribution (see Figure~3 below). It is easy to check that it is correctly normalized and reproduces the leading terms of all the central moments we have computed above.
Of course we have assumed here that the domain of $\mathcal{P}(\hat{q})$ is the whole real line. While this is not strictly true, it is not a problem in the small $\epsilon$ limit we are interested in, since $\hat{q}$ is a rescaled variable, and in terms of the original variable $q$ the distribution looks highly compressed (almost like a $\delta$-function). The distribution falls off exponentially in the positive direction, and even faster (as a double exponential) in the negative direction, so any tails outside the original domain of $\mathcal{P}(q)$ are highly suppressed for small $\epsilon$.
\section{Triple Overlap} \label{Triple}
We will now turn to a finer probe of the structure of the space of states, namely the triple overlap distribution. Here we consider three copies (replicas) of the system under consideration, and compute the multivariate distribution of mutual overlaps $q_{ab}$ between them.
In order to avoid an overabundance of indices we will use dual variables $q_a \equiv |\varepsilon_{abc}|\, q_{bc}/2$, and similarly for other quantities.
\subsection{Moments in Terms of Wigner Functions}
We can express the moments of this distribution in terms of coarse-grained Wigner functions as follows:
\[
\langle q_{1}^{K_{1}} q_{2}^{K_{2}}q_{3}^{K_{3}} \rangle &=& \int_0^{2\pi} \prod_{i=0}^K {\mathrm{d} \theta_i \over 2 \pi} \int \prod_{j=1}^K \mathrm{d} X_j \mathrm{d} P_j \\
&\times& \left(W_1^{K_{1}+1, K_{1}+2, \ldots ,K} \right) \left(W_2^{1,2, \ldots, K_{1}, K_{1}+K_{2}+1, K_{1}+K_{2}+2 \ldots ,K} \right)
\left(W_3^{1,2, \ldots, K_{1}+K_{2}} \right) \ , \nonumber
\]
where there are three groups of degrees of freedom\footnote{Note that the $K_a$ are dual variables corresponding to the more natural $K_{ab}$, whereas the Wigner distributions have only one index to begin with since they are associated with one replica only.} such that $K = K_{1}+K_{2}+K_{3}$. For simplicity we have chosen a particular ordering of the $K$ degrees of freedom, which makes very explicit that every set of variables $(\theta_i,X_i,P_i)$ for a given index $i$ is associated with exactly two of the three replicas (but of course any permutation would be just as good). Reconstructing the full triple overlap distribution from its moments will allow us to determine, amongst other features of the space of states, to what extent the system exhibits ultrametricity.
We can use the representation (\ref{Wigner}) for the Wigner functionals, where the parameters $\lambda_k$ and $\tilde{\lambda}_k$ should be thought of as carrying an additional index to indicate which replica they belong to. As above, however, performing the integral over the coarse-grained expectation values $X_i$ and corresponding momenta $P_i$ leads to delta functions that allow us to eliminate all but one complete set of $(\lambda_k, \tilde{\lambda}_k)$ parameters, such that the extra index essentially disappears.
To illustrate this in more detail, let's assume that the momentum integral again only contributes terms suppressed by positive powers of $\epsilon$, and rescale the $\lambda_k$ as above, so that up to numerical factors the Wigner distribution effectively can be written as
\[
W_a^{1,2,3\ldots K} \sim \int_{-\infty}^\infty \prod_{i = 1}^K \left(\mathrm{d} \lambda_i^a\, e^{-(\lambda_i^a)^2/2}\, e^{2 \pi i \lambda_i^a X_i'}\right) \prod_{k>j\ge1}^K \left(4 \sin^2 { \theta_k-\theta_j \over 2}\right)^{-\lambda_j^a \lambda_k^a / (2\log\epsilon)} \ .
\]
Multiplying three such factors and integrating over the (for simplicity also rescaled) $X_i'$ will set sums of two $\lambda_i^a$ with the same $i$ but different replica index equal to zero. Integrating over those delta functions, and choosing the remaining set of parameters $\lambda_i$ in a symmetric fashion, the resulting formula for the moments of the triple overlap distribution reduces to basically the same expression (\ref{HigherInt}) we have already encountered.
The overall $K$ is now interpreted as the sum of the $K_a$, and the left-over parameters $\lambda_i$ are integrated over with Gaussian weights as before, except for one crucial difference: the angular factors connecting two degrees of freedom in different groups have an extra factor of $-{1 \over 2}$ in the exponent. This is clear, since as shown in Figure~2, those factors appear only in one of the three Wigner distributions (rather than in two of them such as those connecting degrees of freedom in the same group). The most obvious choice of parameters is to define e.g. $\lambda_i \equiv \lambda_i^2 = - \lambda_i^3$ for $i = 1,2,\ldots,K_1$, which introduces minus signs whenever these is a connection between the first and the second group. Similarity transformations, leading to less symmetric choices of parameters, can eliminate some, but not all of these minus signs.
To put it differently, if we carry out the Gaussian $\lambda_i$ integrals first and think of the problem as computing the (inverse square root of the) determinant of a $K$ by $K$ matrix $M_{ij}$, the crucial difference in the structure of the relevant matrix compared to the $M_{ij}$ considered above is that block off-diagonal elements are multiplied by $-{1 \over 2}$, whereas entries in the three $K_a$ by $K_a$ blocks on the diagonal are the same as before. This is very reminiscent of the form of the overlap matrix in a system with one-step replica symmetry breaking (even though there appears to be no direct physical relation).
\begin{figure}
\begin{center}
\vspace{-1cm}
\includegraphics[width = 0.9 \textwidth]{TripleOL}
\vspace{-0.1cm}
\caption{\small Illustration of the triple overlap computation. On the left the three copies of the Wigner distribution are shown, with all edge factors that appear in them. One the right, we superimpose the three pictures, which corresponds to the situation after integrating over the coarse-grained variables $(X_i, P_i)$ and eliminating all $(\lambda_i,\tilde{\lambda}_i)$ parameters that appear in $\delta$-functions. Edges within one group appear twice on the left, and after being identified are drawn as thick lines on the right. Edges that stretch between two groups appear only once, and are suppressed by a factor of $-{1 \over 2}$.}
\vspace{-0.5cm}
\end{center}
\end{figure}
\subsection{Computing Moments Using Graphical Techniques}
Since the expressions for the moments of the distribution are so similar, the calculations of Section \ref{Low} translate immediately into results for the first few moments of the triple overlap distribution. This will not take us very far however, so we'll turn directly to applying the method of Section \ref{Graphs} to this problem.
We have learned that we should really be computing the central moments
\[
\langle \hat{q}_{1}^{K_{1}} \hat{q}_{2}^{K_{2}} \hat{q}_{3}^{K_{3}} \rangle \equiv (-256 \pi^4 \log^3 \epsilon)^{K/2} \left\langle (q_{1} - \langle q_1 \rangle)^{K_{1}} (q_{2}- \langle q_2 \rangle)^{K_{2}}(q_{3}- \langle q_3 \rangle)^{K_{3}} \right\rangle \ .
\]
As above, we will associate to each of these central moments a set of graphs, with $K$ vertices labelled by the index $i$, and edges corresponding to the angular factors connecting two vertices (i.e.~depending on the difference of two degrees of freedom). However, now there is additional structure to the problem: the vertices are split up into three mutually exclusive groups, with the number in each group given by $K_a$, and the edges come in two varieties, namely those connecting two vertices within one group, and those connecting different groups.
The arguments of Subsection \ref{EdgeExp} go through as before, since they don't rely on the exact form of the edge factors, and thus we can write the leading term of each central moment as a sum of terms labelled by graphs with $K$ non-isolated vertices on which minimal length links can be drawn.
The contribution of each cycle or dimer contained within such a graph also follows easily from the calculation in Subsection \ref{CycleGr}. We simply need to insert appropriate factors of $-{1 \over 2}$ whenever we cross the boundary between different groups of vertices.
However, we face a rather nontrivial counting problem: how many (unoriented) graphs with $K$ vertices are there which consist of disconnected cycle or dimer components (subgraphs), in which each vertex is part of exactly one disconnect component, and also belongs to exactly one of three distinguishable groups with $K_a$ members, such that the boundaries between groups are crossed a given number of times?
\subsection{More Graph Combinatorics}
Of course we are not really interested in the number of graphs as such, but rather would like to compute a weighted sum over all of them with the appropriate factors attached to each one, and thus we need a complete classification of the graphs that contribute to the central moments at leading order. We will do this in several steps.
At the highest level, the relevant graphs are characterized by the family they belong to. As discussed in the previous section, we specify the family by an integer partition of K written as $m_L \in \mathsf{P}(2 , K)$ (which each summand at least two), which counts the multiplicities of loops (or dimers) of length $L$. We know from the above that within each family there are $K!\, 2^{m_2} \prod_{L \ge 2} (m_L!)^{-1} (2L)^{-m_L}$ distinct graphs. We obtained this by first counting the distinguishable ways of assigning each of the $K$ vertices to one of the loops, and then multiplying by the number of ways we can string together (wire up) a given set of $L$ vertices into a loop of length $L$.
For the present problem however, we need a finer classification, since vertices are now assigned to one of three groups and we need to consider whether edges cross between different groups. Therefore, when assigning vertices to loops of a certain length $L$ we have to consider how many vertices to take from each of the three groups. The number of ways of doing this is given by
\[
&&{1 \over m_L!} \sum_{n_1^1,n_2^1,n_3^1 = 0}{K_1 \choose n_1^1} {K_2 \choose n_2^1} {K_3 \choose n_3^1} \delta(n_1^1+n_2^1+n_3^1-L) \times \ldots \nonumber \\
&\times& \sum_{n_1^{m_L},n_2^{m_L},n_3^{m_L} = 0}{K_1 \choose n_1^{m_L}} {K_2 \choose n_2^{m_L}} {K_3 \choose n_3^{m_L}} \delta(n_1^{m_L}+n_2^{m_L}+n_3^{m_L}-L) \nonumber \\
&=&{1 \over m_L!} \sum_{n_1^1\ldots n_1^{m_L} = 0} \sum_{n_2^1\ldots n_2^{m_L} = 0} {K_1!K_2!K_3! \over n^1_1!\, n^1_2!\, (L-n^1_1-n^1_2)!\ldots n^{m_L}_1!\,n^{m_L}_2!\,(L-n^{m_L}_1-n^{m_L}_2)! } \nonumber \\
&\times& {1 \over (K_1 - \sum_{\mu=1}^{m_L} n_1^\mu)!(K_2 - \sum_{\nu=1}^{m_L} n_2^\nu)!(K_3 - \sum_{\rho=1}^{m_L} (L-n_1^\rho-n_2^\rho))!} \ .
\]
The factor of $1/m_L!$ appears because loops of the same length are interchangeable at this level. For each loop there is obviously a constraint that the numbers of the vertices drawn from the three groups have to add up to $L$. When we go through all the possible values of $L$ we will multiply factors of this type until all vertices are assigned. Thus, when we are given a set of integers $\{K_1,K_2,K_3\}$ there are also global constraints which impose that the total number of vertices drawn from each group has to add up to the correct $K_a$. Demanding this, the total number of possible vertex assignments is
\[
\sum_{\{n_1^{L,\mu_L}\} = 0}^\infty\, \sum_{\{n_2^{L,\mu_L}\} = 0}^\infty && K_1! K_2! K_3!\, \delta(K_1 - \sum_{L, \nu_L} n_1^{L,\nu_L})\, \delta(K_2 - \sum_{L, \nu_L} n_2^{L,\nu_L}) \nonumber \\
&& \prod_{L \ge 2} {1 \over m_L!} \prod_{\mu_L=1}^{m_L} \,{1 \over n^{L, \mu_L}_1!\,n^{L, \mu_L}_2!\,(L-n^{L, \mu_L}_1-n^{L, \mu_L}_2)!} \ ,
\]
where there are parameters $n^{L,\mu_L}$ appearing inside the sums (and products) for every $L$ for which $m_L >0$ and $\mu_L = 1,2 \ldots m_L$.
Thus for given $K_a$ we can define a genus (think biology, not topology) of graphs belonging to a given family $m_L$ by a set of pairs of integers $\{n_1^{L,\mu_L}, n_2^{L,\mu_L} \}$ where pairs associated with the same length $L$ but different labels $\mu_L$ are interchangeable (i.e.~sets differing only by swapping pairs with the same $L$ are considered equivalent). A valid set has to satisfy the local constraints $n_1^{L,\mu_L} + n_2^{L,\mu_L} \le L$ for all $L$ and $\mu_L$ and the global constraints $\sum_{L,\mu_L} n_a^{L,\mu_L} = K_a$.
However, even within a given genus there are graphs with different factors associated with them, because even once we have assigned vertices to loops of given length there are still different ways of wiring up the vertices into a cycle, which may cross the boundaries between groups a different number of times. As a simple example, consider a cycle of length four for which we have chosen two vertices from one group and two vertices from another. Clearly, we can draw loops of length four that cross between the groups either twice or four times (as is apparent from the three leftmost graphs in Figure~1, considering say the top two vertices as belonging to a different group than the bottom two).
Thus within a given family and genus, we can think of a species of graphs as a subset which for each cycle it contains has a specified number of edges between vertices belonging to different groups, each of which picks up a factor if $-{1 \over 2}$ relative to edges within one group. We define for each cycle a loop factor $\mathcal{L}^L(n_1, n_2)$ that sums over all possible cycle graphs within a given genus and weighs each species by the appropriate power of $-{1 \over 2}$ (taking into account the correct population of each species). In the example of the previous paragraph, we can draw two different graphs that cross boundaries between groups twice and one graph that has four crossings. Therefore for this loop the correct factor would be $\mathcal{L}^4(n_1=2, n_2=2) = 2 (-{1 \over 2})^2 + 1 (-{1 \over 2})^4 = {9 \over 16}$.
The total loop factor will simply be a product of such factors for each cycle (or dimer) contained in the graph.
In summary, in order to compute a given moment with exponents $K_a$ all we need to do is sum over all relevant families, genera and species of graphs with zeta function values depending on the family, and loop factors taking care of the numerical suppression associated with edges crossing borders as well as multiplicities within each species. To execute this we would have to overcome two difficulties: firstly we would need an explicit formula for the loop factors $\mathcal{L}^L(n_1, n_2)$, and secondly we would have to find an efficient way of implementing the global constraints when summing over genera. Fortunately, it is not necessary to perform either of these steps explicitly.
\subsection{Characteristic Function from Matrix Powers}
An elegant way to avoid the ugly details that arise when computing individual moments in this fashion is to work with generating functions. Consider the matrix
\[
\Omega = \left( \begin{array}{ccc}
\omega_1 & 0 & 0 \\
0 & \omega_2 & 0 \\
0 & 0 & \omega_3 \end{array} \right)
\left( \begin{array}{ccc}
1 & -1/2 & -1/2 \\
-1/2 & 1 & -1/2 \\
-1/2 & -1/2 & 1 \end{array} \right) =
\left( \begin{array}{ccc}
\omega_1 & -\omega_1/2 & -\omega_1/2 \\
-\omega_2/2 & \omega_2 & -\omega_2/2 \\
-\omega_3/2 & -\omega_3/2 & \omega_3 \end{array} \right) \ .
\]
It turns out that it generates precisely the loop factors we require
\[ \label{Omega}
{1 + \delta_{L,2} \over 2L}\, \mathrm{tr}(\Omega^L) =\sum_{n_1+n_2+n_3 = L} {\mathcal{L}^L(n_1, n_2) \over n_1!\, n_2!\, n_3!}\ \omega_1^{n_1}\omega_2^{n_2}\omega_3^{n_3} \ .
\]
Intuitively, the matrix powers of $\Omega$ encode the fact that there is a penalty for going from on group to another (in the off-diagonal elements of $\Omega$), while at the same time taking into account all the different paths that visit each vertex once. The trace makes sure that the path closes in the end. Again, the dimer case is special and gives rise to an extra factor of two.
The above expression, which gives us a linear combination of all loop factors for a given $L$, rather than just a particular one, is precisely what we need to compute the characteristic function of the triple overlap distribution. This also avoids the second problem of having to impose global constraints: since we have to sum over all $K_a$, the issue no longer arises.
Neglecting all subleading terms (suppressed by additional inverse powers of $\log\epsilon$) the characteristic function is then given by
\[
&&\langle \exp(i \omega_1 \hat{q}_1 + i \omega_2 \hat{q}_2 + i \omega_3 \hat{q}_3)\rangle = \sum_{K_1, K_2, K_3 = 0}^\infty {(i \omega_1)^{K_1}(i \omega_2)^{K_2}(i \omega_3)^{K_3} \over K_1!\, K_2!\, K_3!} \langle \hat{q}_{1}^{K_{1}} \hat{q}_{2}^{K_{2}} \hat{q}_{3}^{K_{3}} \rangle \nonumber \\
&=& \sum_{K_1, K_2, K_3 = 0}^\infty \sum_{m_L \in \mathsf{P}(2 , K)} \sum_{\{n_1^{L,\mu_L}\} = 0}^\infty \sum_{\{n_2^{L,\mu_L}\} = 0}^\infty \sum_{\{n_3^{L,\mu_L}\} = 0}^\infty \left(\prod_{a=1}^3 \delta(K_a - \sum_{L, \nu_L} n_a^{L,\nu_L}) \right) \nonumber \\
&&\times (i \omega_1)^{K_1}(i \omega_2)^{K_2}(i \omega_3)^{K_3} \prod_{L \ge 2} {1 \over m_L!} \prod_{\mu_L =1}^{m_L} {\mathcal{L}^L(n_1, n_2) \over n_1!\, n_2!\, n_3!} \, \delta(n_1^{L,\mu_L}+n_2^{L,\mu_L}+n_3^{L,\mu_L}-L) \ 2 \zeta(L) \ . \nonumber \\
&=& \exp\left(\sum_{L \ge 2} {i^L \zeta(L) \over L} \, \mathrm{tr}(\Omega^L) \right) \ .
\]
Again, the extra factors of two for the dimer cancel, and performing the sums over the $K_a$ first, then summing over the $n_a^{L,\mu_L}$ using (\ref{Omega}) and finally using the generating function for integer partitions, the expression simplifies dramatically to a non-Abelian version of (\ref{character}).
\subsection{A Tale of Two Gumbels}
The non-zero eigenvalues of $\Omega$ are given by
\[
\omega_\pm = {1\over 2}(\omega_1+\omega_2+\omega_3) \pm {1 \over 2 } \sqrt{\omega_1^2+\omega_2^2+\omega_3^2-\omega_1\omega_2-\omega_1\omega_3-\omega_2\omega_3} \ ,
\]
and therefore the trace of $\Omega^L$ is simply equal to $\omega_+^L + \omega_-^L$.
If we introduce cylindrical polar coordinates aligned with the equilateral axis
\[
\omega_z &=& {1 \over \sqrt{3}} (\omega_1+\omega_2+\omega_3) \ , \nonumber \\
\omega_r &=& \sqrt{2 \over 3} \sqrt{\omega_1^2+\omega_2^2+\omega_3^2-\omega_1\omega_2-\omega_1\omega_3-\omega_2\omega_3} \ , \nonumber \\
\omega_\varphi &=& \arctan\left({\sqrt{3}(\omega_1 - \omega_3) \over \omega_1 - 2 \omega_2 + \omega_3}\right) \ ,
\]
and similarly for the dual coordinates $(\hat{q}_z,\hat{q}_r,\hat{q}_\varphi)$, we find that the characteristic function can be written as
\[ \label{TODChar}
\left\langle e^{i \omega_z \hat{q}_z + i \omega_r \hat{q}_r \cos(\omega_\varphi - \hat{q}_\varphi) } \right\rangle = e^{- i \gamma \sqrt{3}\, \omega_z} \Gamma(1 - i \omega_+) \Gamma(1 - i \omega_-) \ ,
\]
where we have performed the sum over $L$ explicitly, and
$\omega_\pm = (\sqrt{3}/ 2) ( \omega_z \pm \omega_r/\sqrt{2})$.
The characteristic function of the triple overlap distribution is independent of the angular variable $\omega_\varphi$. In fact, it is simply a product of the characteristic functions of two Gumbel distributions, and thus its Fourier transform will lead to an expression resembling a convolution (though not exactly, because the $\omega_\pm$ integration region is only a half-plane). Performing this Fourier transform we obtain, using again the integral representation of the $\Gamma$-function
\[
\mathcal{P}(\hat{q}_1, \hat{q}_2, \hat{q}_3) &=& {1 \over (2 \pi)^3} \int \mathrm{d} \omega_z \mathrm{d} \omega_r \mathrm{d} \omega_\varphi \ \omega_r \, e^{-i \omega.\hat{q}} \, e^{-i \sqrt{3} \gamma \omega_z} \\
&& \int_0^\infty \mathrm{d} s \, e^{-s} \int_0^\infty \mathrm{d} t \, e^{-t} \, e^{-i \sqrt{3}(\log s + \log t)\, \omega_z / 2 } \, e^{-i \sqrt{3}(\log s - \log t)\, \omega_r / (2 \sqrt{2}) } \ . \nonumber
\]
Since the integrand is independent of $\omega_\varphi$ except for the complex exponential of $\omega.\hat{q}$, the $\omega_\varphi$ integral will simply pick out the lowest order Bessel function of the angular expansion of this factor, and the result will be independent of $\hat{q}_\varphi$. The $\omega_z$ integral leads to a delta function which trivializes one of the parametric integrals. After an appropriate change of variables, the second parametric integral then takes the form of the integral representation of the modified Bessel function of the second kind
\[
K_\nu(x) = \int_0^\infty \mathrm{d} u \, e^{-x \cosh u} \cosh(\nu u) \ ,
\]
but with purely imaginary order $\nu$. This leads to the following expression for the triple overlap distribution
\[ \label{TripleDis}
\mathcal{P}(\hat{q}_z, \hat{q}_r) = {2 \over \pi \sqrt{3}} \int_0^\infty \mathrm{d} \omega_r \, \omega_r \, J_0(\omega_r \hat{q}_r) \, e^{-2(\gamma + \hat{q}_z / \sqrt{3})} K_{i \sqrt{3}\, \omega_r / \sqrt{2}} \left(2 e^{-(\gamma + \hat{q}_z / \sqrt{3})} \right) \ .
\]
This peculiar type of integral, with the integration variable appearing in the order of the modified Bessel function, is known as a Kontorovich-Lebedev transform (or its inverse, depending on conventions).
One can check that this probability density is correctly normalized, and that the first moment $\langle \hat{q}_z \rangle$ vanishes\footnote{This might not be apparent from the plot of the probability density, which for small $\hat{q}_r$ is clearly peaked in a region of negative $\hat{q}_z$. Recall however that the integration measure contains another factor of $\hat{q}_r$, which mitigates this, and enhances the tail of the distribution which is more pronounced for positive $\hat{q}_z$.} as required, though for the computation of higher moments it is more convenient to directly expand the characteristic function (\ref{TODChar}) instead.
\begin{figure}
\begin{center}
\vspace{-2.8cm}
\includegraphics[width = 0.43 \textwidth]{Gumbel}
\hspace{-1cm}
\includegraphics[width = 0.6 \textwidth]{DeSitterOverlapTOD}
\vspace{-1cm}
\caption{\small Overlap distribution $\mathcal{P}(\hat{q})$ (left), and triple overlap distribution $\mathcal{P}(\hat{q}_z, \hat{q}_r)$ (right).} \vspace{-0.5cm}
\end{center}
\end{figure}
\section{Discussion}
We have introduced a neat graph-based technique that allowed us to explicitly compute the overlap and triple overlap distributions for a coarse-grained, massless scalar field on a 1+1 dimensional de Sitter background by largely combinatorial methods.
What have we learned from these calculations? The obvious answer, which was of course entirely expected, is that in the naive late time $\epsilon \to 0$ limit, in which the number of causally disconnected boxes in our model universe diverges, the overlap of a scalar field in $dS_2$ tends to vanish (i.e.~the limiting distribution is a delta function centered on zero).
This simply confirms our intuition that in this limit the convolution with the box function merely picks out a perfectly localized harmonic oscillator degree of freedom, and since the underlying field theory is free there is no reason to expect anything non-trivial to emerge from such a collection of harmonic oscillators.
However, it is a remarkable fact that if we look more closely, subtract the mean and scale the central moments of the overlap distribution appropriately, we find a rather non-trivial shape for the overlap, which is given by a Gumbel distribution (\ref{Gumbel}). Thus we have obtained a precise characterization of how likely deviations from the above trivial behavior are.
Whether we accept this as evidence for a non-trivial phase structure, depends primarily on whether or not we consider the rescaled distribution physically relevant.
There may be good, physical reasons to perform this shift and rescaling, since in a sense it merely removes the obvious effect of an inflating universe, which is the tendency to dilute overlaps between states on an absolute scale, maybe not unlike the situation in a spin glass to which we keep adding spins according to some rule. We will not speculate on this further, except to note that the rule in this analogy is by no means random; the rescaling of the overlap we had to perform is merely by powers of $-\log\epsilon$, whereas the number of causally disconnected boxes grows as $\epsilon^{-1}$.
Interestingly, the Gumbel density is also the distribution of a much simpler quantity, namely the (regularized and zero mode subtracted) self-overlap $\int \mathrm{d} \theta | \phi(\theta)|^2$, as explained in \cite{AD}.
A priori there was no reason to expect this heuristic notion of overlap to agree with the full phase-space definition (certainly we wouldn't in general), and this should presumably be taken as a further testament to the simplicity of the system we are studying.
We have also obtained an explicit expression for the triple overlap distribution (\ref{TripleDis}) which adds this system to a rather short list of models in the literature whose triple overlaps are known in closed form.
Again, we could consider first the naive $\epsilon \to 0$ limit, in which this distribution approaches a delta function supported on equilateral triangles (with all three $q_a$ equal to each other), whose size tends to zero as $(-\log \epsilon)^{-1/2}$. While this limiting distribution is technically ultrametric, its ultrametricity is of a trivial nature, since as we have discussed above, in the same limit the regular overlap distribution also approaches a delta function, precluding any interesting phase structure. Furthermore, in a non-trivial tree the pairwise path distances between three arbitrary leaves should be allowed to take different values (only the largest two, but not necessarily all three of them should have to be equal).
However, the appropriately shifted and scaled triple overlap distribution is perfectly smooth, as shown in Figure~3, and given by a rather intriguing Kontorovich-Lebedev integral (\ref{TripleDis}). Still, it is invariant under rotations around the equilateral (i.e.~$z$) axis, which implies that isosceles triangles are not particularly preferred. If there were a non-trivial tree structure underlying these vacuum fluctuations, we wouldn't expect the triple overlap distribution to have support everywhere around the equilateral axis, but only on the three half-planes described by $q_1 = q_2 > q_3$ and permutations thereof. Thus while the rescaled triple overlap distribution certainly has a very interesting structure indicative of ergodicity breaking, we do not see evidence for a hierarchical organization of the state space\footnote{The expression (\ref{TODChar}) was subsequently rederived as the triple distance distribution of a different, simpler measure of overlap in \cite{AD}, where it was used to argue in favor of ultrametricity of the model. This was based on the observation that if one restricts to rare events in the tail of the distribution, the conditional probability $\mathcal{P}(\hat{q}_1| \hat{q}_2, \hat{q}_3)$ is peaked in the region where the three $\hat{q}$'s form isosceles triangles. However, while intriguing in its own right, this approximate localization property is clearly much weaker than ultrametricity, whose definition requires a sharp localization on a two-dimensional surface. Furthermore, it is really a property of a particular conditional distribution, rather than of the joint triple overlap distribution, since it depends crucially on the choice of conditions (which amounts to specifying a scheme for comparing isosceles triangles with others that would not be present in an ultrametric distribution).
In this case the conditions imposed explicitly break the rotational symmetry of the joint distribution about the line of equilateral triangles, which maps isosceles triangles to non-isosceles ones. On the other hand, if we looked at a conditional distribution that respects this symmetry, e.g.~by keeping $\hat{q}_z$ (i.e.~the sum of the three sides of the triangle) constant, we would find no such approximate localization.}.
This result might be a limitation forced upon us by the low dimensionality and simple dynamics of our model. Higher dimensional de Sitter spaces may exhibit a richer structure, and we hope that the calculations presented here can be extended, at least at some level of approximation, to these cases. Other possible generalizations include introducing masses and self-interactions, or considering non-scalar fields, and we hope that the study of such systems will reveal further interesting facts about the statistical properties of inflating universes at the largest scales.
\section*{Acknowledgments}
I am very grateful to F.~Denef for introducing me to the concept of overlap distributions and giving me the idea to perform the computations presented in this note. I have enjoyed many fruitful discussions with him and had the pleasure of his collaboration during the early stages of this project. I am also indebted to D.~Anninos for sharing some of his unpublished notes on this topic with me.
Furthermore, I would like to thank I.~Klebanov for useful comments on the manuscript.
This work was partially supported by DOE grant DE-FG02-92ER40697.
|
\section{Introduction}
Recent polarized beam experiments and global QCD analyses suggest that the contribution of the gluon helicity $\Delta G$ to the
spin of the proton is rather small \cite{Boer:2011fh}. This observation, together with the inexorable fact that the quark helicity contribution $\Delta \Sigma$ is also small (less than $30\%$), lead one to suspect that the key to understand the proton spin puzzle is the orbital angular momentum (OAM) of quarks and gluons. However, progress in this direction has been hindered by a number of difficulties in measuring, and even defining the OAM.
So far, the only well--recognized, gauge invariant definition of the quark OAM is the one by Ji \cite{Ji:1996ek}
which can be measured, indirectly, as the difference between a certain moment of the generalized parton distribution and $\Delta \Sigma$. Although generally accepted, this approach may be criticized on the basis that the corresponding operator is not the `canonical' one that satisfies the fundamental commutation relation of the angular momentum operator in quantum mechanics. Efforts to improve upon this point have led Chen {\it et al.} to propose a completely new decomposition scheme of the QCD angular momentum tensor \cite{Chen:2008ag,Chen:2009mr} which has triggered a flurry of activity lately
\cite{Wakamatsu:2010qj,Wakamatsu:2010cb,Cho:2010cw,Chen:2011gn,Leader:2011za,Hatta:2011zs,Wakamatsu:2011mb,Chen:2011zzh,Zhang:2011rn,Lin:2011us}.
However, the issue still remains very controversial, and the overarching impact of this new formalism as well as its practical usefulness in phenomenology are yet to be clarified.
In this work, we investigate the quark OAM along the line of our previous work \cite{Hatta:2011zs} which we view as the proper rendition of the formalism \cite{Chen:2008ag,Chen:2009mr} in the context of high energy QCD. We shall show that one can represent the canonical OAM as the matrix element of a manifestly gauge invariant operator which turns out to be equivalent to that obtained in the Wigner distribution approach \cite{Lorce:2011kd}. This paves the way to measure the canonical OAM experimentally or numerically on a lattice, and thus helps to mitigate the criticism that the matrix elements defined in \cite{Chen:2008ag} do not have known physical measurements \cite{Ji:2010zza}. Actually, in the gluon helicity sector Ref.~\cite{Hatta:2011zs} has already shown how one can reconcile the gluon helicity defined in \cite{Chen:2008ag} with $\Delta G$ which is measurable.
We now extend this finding to the OAM sector.
\section{Decomposition of the QCD angular momentum operator }
The main idea of \cite{Chen:2008ag,Chen:2009mr} is that one can achieve a complete, gauge invariant decomposition of the QCD angular momentum operator by identifying the `physical' and `pure gauge' components of the gauge field
\begin{eqnarray}
&& A^\mu = A^\mu_{\scriptsize\mbox{phys}} + A^\mu_{\scriptsize\mbox{pure}}\,, \\
&& F_{\scriptsize\mbox{pure}}^{\mu\nu}
= \partial^\mu A^\nu_{\scriptsize\mbox{pure}} -\partial^\nu A^\mu_{\scriptsize\mbox{pure}} +ig [A^\mu_{\scriptsize\mbox{pure}},A^\nu_{\scriptsize\mbox{pure}}] = 0\,, \label{pure}
\end{eqnarray}
which transform differently under gauge transformations
\begin{eqnarray}
&& A^\mu_{\scriptsize\mbox{phys}} \to U^\dagger A^\mu_{\scriptsize\mbox{phys}}U\,, \nonumber \\
&& A^\mu_{\scriptsize\mbox{pure}} \to U^\dagger A^\mu_{\scriptsize\mbox{pure}} U -\frac{i}{g} U^\dagger \partial^\mu U\,. \label{tran}
\end{eqnarray}
The QCD angular momentum tensor $M^{\mu\nu\lambda}$ can then be written as a sum of the helicity and the orbital angular momentum of quarks and gluons. In the `covariant' form \cite{Wakamatsu:2010qj,Wakamatsu:2010cb}
useful for high energy experiments, the original proposal by Chen {\it et al}. \cite{Chen:2008ag,Chen:2009mr} reads
\begin{eqnarray}
M_{\scriptsize\mbox{quark-spin}}^{\mu\nu\lambda} &=& -\frac{1}{2}\epsilon^{\mu\nu\lambda\sigma}\bar{\psi}
\gamma_5 \gamma_\sigma \psi\,, \\
M_{\scriptsize\mbox{quark-orbit}}^{\mu\nu\lambda}&=&\bar{\psi}\gamma^\mu (x^\nu iD_{\scriptsize\mbox{pure}}^\lambda
-x^\lambda iD_{\scriptsize\mbox{pure}}^\nu )\psi\,, \label{24} \\
M_{\scriptsize\mbox{gluon-spin}}^{\mu\nu\lambda}&=& F_a^{\mu\lambda}A_{\scriptsize\mbox{phys}}^{\nu a} -
F_a^{\mu\nu}A_{\scriptsize\mbox{phys}}^{\lambda a} \,, \label{glu} \label{25} \\
M_{\scriptsize\mbox{gluon-orbit}}^{\mu\nu\lambda}&=& F_a^{\mu\alpha}\bigl(x^\nu (D_{\scriptsize\mbox{pure}}^\lambda A_\alpha^{\scriptsize\mbox{phys}})_a
-x^\lambda (D^\nu_{\scriptsize\mbox{pure}}A_\alpha^{\scriptsize\mbox{phys}})_a \bigr)\,. \label{26}
\end{eqnarray}
where $D^\nu_{\scriptsize\mbox{pure}} \equiv \partial^\nu +igA^\nu_{\scriptsize\mbox{pure}}$, and $a,b=1,2,\cdots,8$ are the color indices.\footnote{
Our convention is $
\epsilon^{0123}=+1\,, \gamma_5=-i\gamma^0\gamma^1\gamma^2\gamma^3$.
We shall use the light--cone coordinates $x^\pm = \frac{1}{\sqrt{2}}(x^0\pm x^3)$ and denote the transverse coordinates with latin indices $x_T = \{x^i\}$, $(i,j,\cdots=1,2)$. The two--dimensional antisymmetric tensor is defined as
$\epsilon^{ij}=-\epsilon^{+-ij}\,, \epsilon^{12}=\epsilon_{12}=-\epsilon^{21}=1$.
} Using the transformation rule (\ref{tran}), it is easy to check that each of the above components is gauge invariant. Such a complete decomposition goes beyond Ji's framework in which the gluonic part cannot be separated into the helicity and orbital parts. The price to pay, however, is that the decomposition is not local in the sense that $A^\mu_{\scriptsize\mbox{phys}}$ is in general nonlocally related to the total $A^\mu$. Moreover, it is not entirely covariant, either, because $A^\mu_{\scriptsize\mbox{phys}}$ actually depends on the frame as we shall soon see.
An alternative decomposition of the orbital part was suggested by Wakamatsu \cite{Wakamatsu:2010qj}
\begin{eqnarray}
M_{\scriptsize\mbox{quark-orbit}}^{'\mu\nu\lambda}&=&\bar{\psi}\gamma^\mu (x^\nu iD^\lambda
-x^\lambda iD^\nu )\psi\,, \label{ji} \\
M_{\scriptsize\mbox{gluon-orbit}}^{'\mu\nu\lambda}&=& F_a^{\mu\alpha}\bigl(x^\nu (D_{\scriptsize\mbox{pure}}^\lambda A_\alpha^{\scriptsize\mbox{phys}})_a
-x^\lambda (D^\nu_{\scriptsize\mbox{pure}}A_\alpha^{\scriptsize\mbox{phys}})_a \bigr) \nonumber \\ && \qquad \qquad \qquad \qquad
+ (D_\alpha F^{\alpha\mu})_a
(x^\nu A_{\scriptsize\mbox{phys}}^{\lambda a} -x^\lambda A_{\scriptsize\mbox{phys}}^{\nu a}) \,. \label{waka}
\end{eqnarray}
The second term of (\ref{waka}) is gauge invariant on its own. Using the equation of motion $D_\alpha F^{\alpha \mu}_a = g\bar{\psi}\gamma^\mu t^a \psi$, one sees that it accounts for the difference between $D^\nu_{\scriptsize\mbox{pure}}$ in (\ref{24}) and $D^\nu$ in (\ref{ji}).
There is no consensus as to which definition, (\ref{24}) or (\ref{ji}), is more appropriate for the quark orbital angular momentum. To some extent, it is a matter of choice. It is (\ref{24}), but not (\ref{ji}), that is compatible with the (equal--time) canonical commutation relation of the angular momentum operator $\vec{L}\times \vec{L}=i\vec{L}$,
\begin{eqnarray}
\vec{L}=\vec{x} \times i\vec{D}_{\scriptsize\mbox{pure}}\,, \qquad \vec{x}=(x^1,x^2,x^3)\,.
\end{eqnarray}
The pure gauge condition (\ref{pure}) is crucial for this. (\ref{24}) may thus be called the canonical angular momentum.\footnote{Throughout this paper, we associate the term `canonical' with the operator $iD^\mu_{\scriptsize\mbox{pure}}$ instead of the usual $i\partial^\mu$. The former is actually the gauge covariant generalization of the latter without affecting the commutation relation.}
On the other hand, (\ref{ji}) is the same as Ji's definition \cite{Ji:1996ek} and is accessible from the analysis of the generalized parton distribution (GPD), whereas it has not been known how to measure (\ref{24}).
We will derive an explicit expression of the canonical angular momentum (\ref{24}) in terms of a manifestly gauge invariant operator whose matrix element is, in principle, related to experimental processes or observables in lattice QCD simulations. For this purpose, one must specify what $A^\mu_{\scriptsize\mbox{phys}}$ is. There are several proposals for $A^\mu_{\scriptsize\mbox{phys}}$ in the literature \cite{Chen:2009mr,Cho:2010cw,Hatta:2011zs,Zhang:2011rn}. These definitions are not equivalent as suggested by the work of Ref.~\cite{Chen:2011gn} which showed that they give different values of the gluon helicity (the proton matrix element of (\ref{25})). Here we employ the one proposed in \cite{Hatta:2011zs}
\begin{eqnarray}
A^{\mu }_{\scriptsize\mbox{phys}}(x) = -\int dy^- {\mathcal K}(y-x) {\mathcal W}^{-}_{xy}
F^{+\mu}(y^-,\vec{x}){\mathcal W}^{-}_{yx}\,, \label{hatta}
\end{eqnarray}
where we use the notation $\vec{x}=(x^+,x^i)$ from now on. ${\mathcal W}$ is the Wilson line operator
\begin{eqnarray}
{\mathcal W}^{-}_{xy} \equiv {\mathcal P}
\exp\left(-ig\int^{x^-}_{y^-} A^+(y'^-,\vec{x}) dy'^- \right)\,,
\end{eqnarray}
in the fundamental representation. The superscript `$-$' denotes that the path ordering is in the $x^-$ direction.
${\mathcal K}(y^-)$ is either $\frac{1}{2}\epsilon(y^-)$, $\theta(y^-)$ or $-\theta(-y^-)$, depending on the boundary condition at $x^- = \pm \infty$ in the light--cone gauge $A^+=0$.\footnote{From the viewpoint of the $PT$ (parity and time--reversal) symmetry which will be crucially used below, it seems that the choice ${\mathcal K}(y^-)=\frac{1}{2}\epsilon(y^-)$ is the most natural and convenient one, although the difference does not matter in the end. }
The pure gauge part $A_{\scriptsize\mbox{pure}}$ is
\begin{eqnarray}
A^\mu_{\scriptsize\mbox{pure}}(x)\equiv - \frac{i}{g}{\mathcal W}^{-}_{x,\pm\infty}{\mathcal W}_{\pm\infty} \partial^\mu ({\mathcal W}^{-}_{x,\pm\infty}{\mathcal W}_{\pm\infty})^\dagger\,, \label{pu}
\end{eqnarray}
where ${\mathcal W}_{\pm\infty}={\mathcal P}
\exp\left(-ig\int^{\vec{x}}_{\vec{\infty}} \vec{A}(\pm \infty,\vec{x}')\cdot d\vec{x}' \right)$ is the Wilson line in the spatial direction at $x^- = \pm \infty$.
It represents the residual gauge symmetry of the light--cone gauge $A^+=0$, and is fixed by specifying the boundary condition of the gauge field at $x^- \to \pm \infty$ mentioned above.
It has been shown in \cite{Hatta:2011zs} that (\ref{hatta}) and (\ref{pu}) are a viable decomposition of the total gauge field $A^\mu$. We wish to stress that this particular choice is singled out among others by the criterion of measurability: The corresponding gluon helicity coincides with the usual gluon helicity $\Delta G$ that has been measured in experiments.
Note that the definition (\ref{hatta}) already selects a particular frame---the infinite momentum frame where the partonic interpretation of hadrons is clearest. As emphasized in \cite{Goldman:2011vs}, the decomposition of spin into the helicity and the orbital parts cannot be made entirely covariant, but depends on the frame of reference.
\section{Potential angular momentum}
We now focus on the orbital angular momentum of quarks inside a longitudinally polarized proton. It is given by the forward matrix element of the $\mu\nu\lambda=+ij$ component of (\ref{24}) or (\ref{ji})
\begin{eqnarray}
\epsilon^{ij}L_{\scriptsize\mbox{Chen}}&\equiv & \frac{1}{2P^+}\frac{ \langle PS| \int dx^-d^2x_T M_{\scriptsize\mbox{quark-orbit}}^{+ij}| PS\rangle }{(2\pi)^3\delta^{3}(0) }\nonumber \\
&=& \frac{1}{2P^+} \frac{\langle PS| \int dx^-d^2x_T \, \bar{\psi}\gamma^+ (x^i iD^j_{\scriptsize\mbox{pure}}- x^j iD_{\scriptsize\mbox{pure}}^i)\psi|PS\rangle }{ (2\pi)^3\delta^{3}(0)} \,, \label{jm}
\end{eqnarray}
\begin{eqnarray}
\epsilon^{ij}L_{\scriptsize\mbox{Ji}} &\equiv & \frac{1}{2P^+}\frac{\langle PS| \int dx^- d^2x_T M_{\scriptsize\mbox{quark-orbit}}^{'+ij}| PS\rangle}{(2\pi)^3\delta^{3}(0)} \nonumber \\
&=& \frac{1}{2P^+} \frac{\langle PS| \int dx^- d^2x_T \, \bar{\psi}\gamma^+ (x^i iD^j- x^j iD^i)\psi |PS\rangle}{(2\pi)^3\delta^{3}(0)} \,,
\end{eqnarray}
where $P^2=-S^2=M^2$ (the proton mass squared) and $(2\pi)^3\delta^3(0)=\int dx^- d^2x_T$ is the momentum space delta function. The longitudinal polarization means $S^\mu =(S^+,S^-,S_T) \approx (S^+,0,0_T)$.
We first observe that, since $A^\mu_{\scriptsize\mbox{pure}}=0$ in the light--cone gauge $A^+=0$ \cite{Hatta:2011zs}, $L_{\scriptsize\mbox{Chen}}$ is actually identical to the Jaffe--Manohar (JM) definition \cite{Jaffe:1989jz} of the quark orbital angular momentum
\begin{eqnarray}
L_{\scriptsize\mbox{JM}}=
\frac{1}{2P^+}\frac{\langle PS| \int dx^- d^2x_T\, \bar{\psi}\gamma^+(x^1 i\partial^2-x^2 i\partial^1) \psi |PS\rangle_{\scriptsize\mbox{LC}}}{(2\pi)^3\delta^{3}(0)}\,.
\end{eqnarray}
The subscript ${\scriptsize\mbox{LC}}$ means that the matrix element is evaluated in the light--cone gauge.
In other words, $L_{\scriptsize\mbox{Chen}}$ is the generalization of $L_{\scriptsize\mbox{JM}}$ to arbitrary gauges (see, also, \cite{Wakamatsu:2010qj}). We thus unify the notations $L_{\scriptsize\mbox{Chen}} =L_{\scriptsize\mbox{JM}} \equiv L_{\scriptsize\mbox{can}}$ by introducing the canonical orbital angular momentum $L_{\scriptsize\mbox{can}}$, and write
\begin{eqnarray}
L_{\scriptsize\mbox{Ji}} = L_{\scriptsize\mbox{can}} +L_{\scriptsize\mbox{pot}}\,,
\end{eqnarray}
where the so--called potential angular momentum \cite{Burkardt:2008ua,Wakamatsu:2010qj} is, with our choice of
$A^\mu_{\scriptsize\mbox{phys}}$,
\begin{eqnarray}
\epsilon^{ij}L_{\scriptsize\mbox{pot}} &=&
\frac{1}{2P^+ (2\pi)^3\delta^{3}(0)}\langle PS| \int dx^- d^2x_T\, x^i\bar{\psi}(x) \gamma^+ (-g)(x^iA^j_{\scriptsize\mbox{phys}} -x^jA^i_{\scriptsize\mbox{phys}})\psi(x)|PS\rangle \nonumber \\
&=& \frac{1}{2P^+(2\pi)^3\delta^{3}(0) } \langle PS| \int dx^- d^2x_T\, \Bigl\{ x^i\bar{\psi}(x)
\gamma^+ \int dy^- {\mathcal K}(y^- - x^-) {\mathcal W}^-_{xy}\,
gF^{+j}(y^-,\vec{x}){\mathcal W}^-_{yx} \psi(x)
\nonumber \\
&& - x^j \bar{\psi}(x) \gamma^+ \int dy^- {\mathcal K}(y^- - x^-) {\mathcal W}^-_{xy}\, g
F^{+i}(y^-,\vec{x}){\mathcal W}^-_{yx} \psi(x) \Bigr\} |PS\rangle \,.
\label{phys}
\end{eqnarray}
Now consider the non-forward matrix element
\begin{eqnarray}
\frac{1}{2\bar{P}^+(2\pi)^3\delta^{3}(0)} \langle P'S'|\int dx^- d^2x_T\, x^i\bar{\psi}(x)
\gamma^+
\int dy^- {\mathcal K}(y^- - x^-)\, {\mathcal W}^-_{xy} \, g F^{+j}(y^-,\vec{x}) {\mathcal W}^-_{yx} \psi(x)
|PS\rangle\,, \nonumber
\end{eqnarray}
where $\bar{P}^\mu=(P^\mu+P'^\mu)/2$ and the momentum transfer will be denoted as $\Delta^\mu \equiv P'^\mu-P^\mu$.
The explicit factor $x^i$ can be traded for the
derivative with respect to $\Delta^i$ such that
\begin{eqnarray}
&& \epsilon^{ij}L_{\scriptsize\mbox{pot}} \label{use}
\\
&&= \frac{1}{2P^+}\lim_{\Delta \to 0}
\Bigl\{ \frac{\partial}{i\partial \Delta^i} \langle P'S'| \bar{\psi}(0)
\gamma^+
\int dy^- {\mathcal K}(y^- ) \, {\mathcal W}^-_{0y}\, gF^{+j}(y^-) {\mathcal W}^-_{y0} \psi(0)|PS\rangle
-(i\leftrightarrow j) \Bigr\}\,. \nonumber
\end{eqnarray}
Parity and time--reversal ($PT$) symmetry tells that the following parametrization of the matrix element is possible
\begin{eqnarray}
\langle P'S'| \bar{\psi}(0)
\gamma^+
\int dy^- {\mathcal K}(y^- ) \, {\mathcal W}^-_{0y}\, gF^{+i}(y^-) {\mathcal W}^-_{y0} \psi(0)|PS\rangle
=i\epsilon^{ij} \Delta_j \bar{S}^+ h(\xi) + \cdots\,, \label{th}
\end{eqnarray}
where $\bar{S}=(S+S')/2$ and $\xi \equiv -\Delta^+/2\bar{P}^+$ is the skewness parameter. [The dependence of $h(\xi)$ on the renormalization scale is implicit. We also suppress the dependence on $\Delta^2 \approx -\Delta_T^2$ since it is of higher order. Similar comments apply to other distributions defined below.]
This leads to
\begin{eqnarray}
L_{\scriptsize\mbox{pot}}= h(0)
\frac{S^+}{P^+}\,.
\end{eqnarray}
(\ref{th}) thus defines the potential angular momentum as the matrix element of a manifestly gauge invariant operator.\\
The quark--gluon mixed operator that appears in the matrix element (\ref{th}) is familiar in the context of the twist--three mechanism of the single spin asymmetry (SSA). Let us pursue this analogy and consider the following non-forward matrix element
\begin{eqnarray}
T^{\mu\nu}(x_1,x_2,\xi) &=& \int \frac{dy^- dz^-}{(2\pi)^2} e^{\frac{i}{2}(x_1+x_2)\bar{P}^+z^- + i(x_2-x_1)\bar{P}^+y^-} \nonumber \\
&& \qquad \qquad \times \langle P'S'| \bar{\psi}(-z^-/2)
\gamma^+
{\mathcal W}^-_{\frac{-z}{2}y}\, gF^{\mu\nu}(y^-) {\mathcal W}^-_{y\frac{z}{2}} \psi(z^-/2)|PS\rangle \nonumber \\
&=& \frac{1}{\bar{P}^+}\epsilon^{\mu\nu\rho \sigma}\bar{S}_\rho \bar{P}_\sigma \Psi(x_1,x_2,\xi) + \frac{1}{\bar{P}^+} \epsilon^{\mu\nu\rho\sigma}\bar{S}_\rho \Delta_\sigma \Phi(x_1,x_2,\xi)+ \cdots\,. \label{nonf}
\end{eqnarray}
By symmetry considerations, it follows that $\Psi(x_1,x_2,\xi) = \Psi(x_2,x_1,-\xi)$ and $\Phi(x_1,x_2,\xi)=-\Phi(x_2,x_1,-\xi)$. In the forward limit, and in the transversely polarized case $S^\mu = \delta^\mu_i S^i$, only the $\Psi$--term survives. The function $\Psi(x_1,x_2,0)$ plays the cental role in the so--called soft gluonic pole mechanism of the SSA \cite{Qiu:1991wg,Koike:2006qv}.
In the longitudinally polarized case $\bar{S}^\mu\approx \delta^\mu_+ \bar{S}^+$, the $\Psi$--term vanishes for the relevant component $\mu\nu=+j$.
By performing Fourier transformations, we find
\begin{eqnarray}
&& \langle P'S'| \bar{\psi}(0)
\gamma^+
\int dy^- {\mathcal K}(y^- )\, {\mathcal W}_{0y} \, g F^{+j}(y^-) {\mathcal W}^-_{y0} \psi(0)|PS\rangle
\nonumber \\ &&\qquad = i\bar{P}^+\int dx_1 dx_2\, {\mathcal K}(x_1-x_2) T^{+j}(x_1,x_2) \nonumber \\
&& \qquad = i\epsilon^{jk}\bar{S}^+ \Delta_k \int dXdx\, {\mathcal K}(x) \Phi(X,x,\xi)\,,
\end{eqnarray}
where we switched to the notation $X=\frac{x_1+x_2}{2}$, $x=x_1-x_2$. The kernel is
\begin{eqnarray}
{\mathcal K}(x) = \mbox{p.v.}\frac{1}{x} = \frac{1}{2}\left(\frac{1}{x+i\epsilon} + \frac{1}{x-i\epsilon} \right)\,,
\end{eqnarray}
in the case ${\mathcal K}(y^-)=\frac{1}{2}\epsilon(y^-)$ and
\begin{eqnarray}
{\mathcal K}(x) = \frac{1}{x \pm i\epsilon}\,,
\end{eqnarray}
in the cases ${\mathcal K}(y^-) = \pm \theta(\pm y^-)$.
Comparing with (\ref{th}), we obtain an alternative expression for the potential angular momentum
\begin{eqnarray}
L_{\scriptsize\mbox{pot}}=\int dX dx\, {\mathcal K}(x) \Phi(X,x,0)\,.
\end{eqnarray}
Note that, since $\Phi(X,0,0)=0$, different choices for ${\mathcal K}$ lead to the same result, as they should.
\section{Canonical angular momentum}
Next we exploit the relation between the twist--three approach to the SSA and the approach based on the transverse momentum dependent distribution (TMD).\footnote{The following discussion is similar to the works of Ref.~\cite{Hagler:2003jw,Lorce:2011kd}. We improve upon these works by fully taking into account the gauge field and the issue of gauge invariance.}
In the longitudinally polarized and non-forward case, we define
\begin{eqnarray}
&&f(x,q_T,\Delta) \equiv \int \frac{dz^- d^2z_T}{(2\pi)^3} e^{ix\bar{P}^+z^- -iq_T \cdot z_T}
\label{bou} \\
&& \qquad \times \langle P'S'|\bar{\psi}(-z^-/2,-z_T/2)\gamma^+ {\mathcal W}^-_{\frac{-z}{2},\pm \infty}{\mathcal W}^T_{\frac{-z_T}{2}, \frac{z_T}{2}}{\mathcal W}^-_{\pm \infty, \frac{z}{2}}\psi(z^-/2,z_T/2)|PS\rangle\,, \nonumber
\end{eqnarray}
where ${\mathcal W}^T$ is the Wilson line in the transverse direction at $x^-=\pm \infty$.
In the forward case $\Delta=0$, the matrix element (\ref{bou}) reduces to the usual TMD. As is well--known in that context, there is freedom in choosing the path connecting the points $(-z^-/2,-z_T/2) \to (z^-/2,z_T/2)$. Using the Wilson line that goes to future infinity and then comes back $-z^-/2\to +\infty \to z^-/2$, one takes care of the final state interaction. The TMD in this case is relevant to the semi-inclusive DIS (SIDIS). The other case $-z^-/2 \to -\infty \to z^-/2$ includes the initial state interaction relevant to the Drell--Yan process. For the present purpose, one may as well take the average of the two cases.
The relation between (\ref{bou}) and (\ref{th}) is revealed by taking the second moment of $f$ in $q_T$ \cite{Boer:2003cm}\footnote{Cf. Eq.~(39) of \cite{Boer:2003cm}. Via partial integration, $q_T^i$ is replaced by the spatial derivative $\partial^i_T$. When acting on the Wilson line, it brings down the factor $\partial^i_T A^+ =D^+A^i+F^{i+}$ which reduces to the two terms in (\ref{vani}).}
\begin{eqnarray}
&& F^i(x,\Delta) \equiv \int d^2q_T\, q_T^i f(x,q_T,\Delta) \nonumber \\
&& \qquad
=\frac{1}{2}\int \frac{dz^-}{2\pi}e^{ix\bar{P}^+z^-} \Biggl\{ \langle P'S'|\bar{\psi}(-z^-/2)\gamma^+ \left({\mathcal W}^-_{\frac{-z}{2},\frac{z}{2}} i\overrightarrow{D}^i - i\overleftarrow{D}^i{\mathcal W}^-_{\frac{-z}{2},\frac{z}{2}}\right)\psi(z^-/2)|PS\rangle
\nonumber \\
&& \qquad \qquad \qquad \qquad - \langle P'S'| \bar{\psi}(-z^-/2)\gamma^+ \int dy^-\left({\mathcal K}(y^- -z^-/2)+{\mathcal K}(y^- + z^-/2)\right) \nonumber \\
&& \qquad \qquad \qquad \qquad \qquad \qquad \qquad \times {\mathcal W}^-_{\frac{-z}{2},y}\, gF^{+i}(y^-){\mathcal W}^-_{y,\frac{z}{2}}\psi(z^-/2)|PS\rangle \Biggr\}\,, \label{vani}
\end{eqnarray}
where the kernel ${\mathcal K}$ is in one--to--one correspondence with the choice of the Wilson line path in (\ref{bou}). In the forward case, (\ref{vani}) vanishes for the longitudinal polarization by rotational symmetry in the transverse plane. In the non-forward case, however, the following structure
\begin{eqnarray}
f(x,q_T,\Delta)\sim \frac{i}{\bar{P}^+}\epsilon^{+-ij}\bar{S}^+q_{Ti} \Delta_j \tilde{f}(x,q_T^2,\xi,\Delta_T\cdot q_T)\,,
\end{eqnarray}
is allowed, so that (\ref{vani}) is not necessarily zero. Note that, because of an extra minus sign from $\Delta_j$ under the $PT$ transformation, the function $\tilde{f}$ does {\it not} change signs when changing the directions of the Wilson line,\footnote{More precisely, $\tilde{f}(x,q_T^2,\xi,\Delta_T\cdot q_T) \to +\tilde{f}(x,q_T^2,-\xi,-\Delta_T\cdot q_T)$ after changing the directions. But the difference is immaterial in the limit $\Delta \to 0$.} in contrast to the known sign flip of the spin--dependent TMDs in the SIDIS and Drell--Yan reactions \cite{Collins:2002kn}.
We then take the first moment in $x$
\begin{eqnarray}
&& \int dx F^i(x,\Delta) =
\frac{1}{\bar{P}^+} \Bigl\{ \frac{1}{2}\langle P'S'|\bar{\psi}(0)\gamma^+ (i\overrightarrow{D}^i -i\overleftarrow{D}^i)\psi(0)|PS\rangle
\nonumber \\
&& \qquad \qquad \qquad \qquad - \langle P'S'| \bar{\psi}(0)\gamma^+ \int dy^-{\mathcal K}(y^- ){\mathcal W}^-_{0y}\, gF^{+i}(y^-){\mathcal W}^-_{y0}\psi(0)|PS\rangle \Bigr\} \nonumber \\
&& \qquad \qquad = \frac{1}{\bar{P}^+} \Bigl\{ \frac{1}{2}\langle P'S'|\bar{\psi}(0)\gamma^+ (i\overrightarrow{D}^i -i\overleftarrow{D}^i )\psi(0)|PS\rangle
+ \langle P'S'| \bar{\psi}(0)\gamma^+A^i_{\scriptsize\mbox{phys}} \psi(0)|PS\rangle \Bigr\}
\nonumber \\ && \qquad \qquad = \frac{1}{2\bar{P}^+} \langle P'S'|\bar{\psi}(0)\gamma^+ (i\overrightarrow{D}^i_{\scriptsize\mbox{pure}} -i\overleftarrow{D}^i_{\scriptsize\mbox{pure}} )\psi(0)|PS\rangle\,.
\end{eqnarray}
The matrix element (\ref{th}) indeed appears in the second term of the first equality, but somewhat remarkably, it is absorbed by the covariant derivative in the first term. The final expression features precisely the `pure gauge' part of the covariant derivative $D_{\scriptsize\mbox{pure}}$.
Differentiating with respect to $\Delta$, we arrive at
\begin{eqnarray}
\epsilon^{ij}L_{\scriptsize\mbox{can}} &=& \frac{1}{2P^+}\lim_{\Delta \to 0} \frac{\partial}{i\partial \Delta^i }\langle P'S'|\bar{\psi}(0)\gamma^+ (i\overrightarrow{D}^j_{\scriptsize\mbox{pure}} -i\overleftarrow{D}^j_{\scriptsize\mbox{pure}})\psi(0)|PS\rangle \nonumber \\
&=& \lim_{\Delta \to 0} \frac{\partial}{i\partial \Delta^i } \int dx F^j(x,\Delta)
\nonumber \\
&=& \lim_{\Delta \to 0} \frac{\partial}{i\partial \Delta^i } \int dx d^2q_T\,q_T^j f(x,q_T,\Delta) \nonumber \\
&=& \epsilon^{ij}\frac{S^+}{P^+} \frac{1}{2}\int dx d^2q_T\, q_T^2 \tilde{f}(x,q_T^2)\,.
\label{canoni}
\end{eqnarray}
This is a formula relating the canonical OAM of quarks to the matrix element of a well--defined, manifestly gauge invariant operator.\footnote{We note that a possible UV regularization of the $q_T$--integral (\ref{canoni}) and its evolution require an entirely separate analysis.} The final expression in (\ref{canoni}) agrees with
the OAM constructed by Lorc\'e and Pasquini \cite{Lorce:2011kd} from the Wigner distribution (neglecting gauge invariance). We have thus established a gauge invariant link between the Wigner distribution approach and the spin decomposition framework of Chen {\it et al.}
Similarly, for the gluon orbital angular momentum one can define
\begin{eqnarray}
&&g(x,q_T,\Delta) \equiv -i \int \frac{dz^- d^2z_T}{(2\pi)^3} e^{ix\bar{P}^+z^- -iq_T \cdot z_T}
\label{oamg} \\
&& \qquad \quad \times \langle P'S'|F^{+\alpha}(-z^-/2,-z_T/2) {\mathcal W}^-_{\frac{-z}{2},\pm \infty}{\mathcal W}^T_{\frac{-z_T}{2}, \frac{z_T}{2}}{\mathcal W}^-_{\pm \infty, \frac{z}{2}}A_{\alpha}^{\scriptsize\mbox{phys}}(z^-/2,z_T/2)|PS\rangle\,, \nonumber
\end{eqnarray}
where $A_{\scriptsize\mbox{phys}}$ is as in (\ref{hatta}), and now the Wilson lines are in the adjoint representation. Under the $PT$ transformation, one gets the same operator back (up to the direction of the Wilson line) provided the skewness parameter $\xi$ vanishes, which we assume here. In deeply virtual Compton scattering (DVCS), $\xi=0$ corresponds to the elastic scattering of the photon.
Proceeding as before, one finds the double moment in $x$ and $q_T$
\begin{eqnarray}
\int dx d^2q_T\, q_T^i g(x,q_T,\Delta) = \frac{1}{2\bar{P}^+} \langle P'S'|F^{+\alpha} (\overrightarrow{D}^i_{\scriptsize\mbox{pure}} -\overleftarrow{D}^i_{\scriptsize\mbox{pure}} )A_{\alpha}^{\scriptsize\mbox{phys}}|PS\rangle\,.
\end{eqnarray}
The matrix element of (\ref{26}) thus becomes
\begin{eqnarray}
\frac{1}{2P^+} \frac{\langle PS| \int dx^- d^2x_T\, M_{\scriptsize\mbox{gluon-orbit}}^{+ij}|PS\rangle}{(2\pi)^3\delta^3(0)}&=& \lim_{\Delta \to 0} \frac{\partial}{i\partial \Delta^i } \int dx d^2q_T\,q_T^j g(x,q_T,\Delta) \nonumber \\
&=& \epsilon^{ij} \frac{S^+}{P^+}\frac{1}{2}\int dx d^2q_T \, q_T^2 \tilde{g}(x,q_T^2) \,, \label{glu}
\end{eqnarray}
where we parameterized, as $\Delta \to 0$,
\begin{eqnarray}
g(x,q_T,\Delta)= \frac{i}{\bar{P}^+}\epsilon^{+-ij}\bar{S}^+q_{Ti} \Delta_j \tilde{g}(x,q_T^2) + \cdots\,.
\end{eqnarray}
\section{Conclusions}
We have shown that the physical part of the gauge field $A_{\scriptsize\mbox{phys}}$ proposed in \cite{Hatta:2011zs} leads to well--defined expressions for the canonical and potential angular momenta in terms of the matrix element of certain gauge invariant operators. If one defines $\Delta G$ as the gluon helicity, consistency requires that (\ref{canoni}) is the corresponding canonical angular momentum. Now that we can, at least in principle, measure $L_{\scriptsize\mbox{Ji}}$, $L_{\scriptsize\mbox{can}}$ and the difference $L_{\scriptsize\mbox{pot}}=L_{\scriptsize\mbox{Ji}}-L_{\scriptsize\mbox{can}}$ separately,
it seems more legitimate to call $L_{\scriptsize\mbox{can}}$, or equivalently the OAM from the Wigner distribution \cite{Lorce:2011kd,Lorce':2011ni}, the quarks' genuine orbital angular momentum since it satisfies the fundamental commutation relation.
Regarding measurability, it should be possible to compute (\ref{canoni}) in lattice QCD simulations as in the case of the ordinary (forward) TMD (see, e.g., \cite{Musch:2010ka}).
Since there is no sign flip in (\ref{bou}) when changing the directions of the Wilson lines, the matrix element may not be very sensitive to the choice of the path. If so, and if one is not interested in the $x$--dependence, one may first integrate over $x$ in (\ref{bou}) and connect the quark operators (at the same value of $z^-=0$) by a purely spatial Wilson line in the transverse plane. This avoids the introduction of lightlike Wilson lines on a Euclidean lattice which appears to be a vague issue. For the gluon orbital angular momentum (\ref{glu}) with $g$ defined in (\ref{oamg}), one has to deal with lightlike Wilson lines even after the $x$--integration.
Finally, from the experimental point of view, the matrix elements such as (\ref{nonf}) are related to the twist--three GPDs \cite{Anikin:2000em,Belitsky:2000vx} which are hard to extract. It would be interesting to see if there are processes in which these functions contribute to the cross section at leading order as in the single spin asymmetry.\\
{\it Acknowledgments}---I am indebted to Kazuhiro Tanaka for many discussions, and for reading the manuscript and bringing \cite{Hagler:2003jw} to my attention. I also thank Terry Goldman, Yuji Koike and Jian-Wei Qiu for helpful conversations and correspondence.
This work is supported by Special Coordination Funds for Promoting Science and Technology
of the Ministry of Education, the Japanese Government.
|
\section{Introduction}
Flares are one of the most important components of solar activity and
result in sudden releases of magnetic energy and in massive magnetic
field restructuring in the host active regions. They are made of two
distinct phases: a very short impulsive phase - when magnetic energy is
suddenly released and plasma is rapidly heated to temperatures that may
exceed 15~MK - and a long decay phase, where the plasma cools back to
lower temperatures. Most of flare observations have been carried out
using crystal spectrometers
observing below 20~\AA\ or narrow-band imagers with bandpasses optimized
at wavelengths below 50~\AA, where the emission from multimillion degrees
plasmas is best observed. X-ray spectral observations of the decay phase
of flares have invarably shown a slow decrease of the plasma temperature,
down to a few million degrees. A large body of literature has been devoted
to study the properties of such decaying plasmas. When the flare plasma
temperature decreases below a few million degrees, it fades from X-ray
instruments and cannot be studied anymore with X-ray spectroscopic diagnostic
techniques. As a consequence, no diagnostic studies could be made of
the evolution of the post-flare plasmas in the last phases of cooling.
EUV imagers, such as SOHO/EIT, TRACE, STEREO/EUVI and SDO/AIA, sample quiescent
coronal temperatures of $\approx 1-3$~MK and thus can extend decay phase
flare observations down to lower temperatures. However, since they rely
on broad-band filters, they can only provide limited diagnostic results.
SUMER, observing in the 500-1600~\AA\ UV range, provides a golden opportunity
to monitor the behavior of decaying plasma down to chromospheric temperatures,
thanks to the multitude of coronal, transition region and chromospheric lines
that populate its spectral range. During its operational life, SUMER
observed a number of C and M flares,
and some of its observations lasted enough to let the post-flare plasma cover
the entire evolution from the impulsive to the very end of the decay phase.
However, the spectroscopic diagnostic potential of SUMER has never been
exploited to try to study the evolution of the post-flare plasma below 2~MK.
The aim of the present work is to carry out such a study.
Recently, Feldman {\em et al.\ } (2003) reported on SUMER observations of the evolution
a C8 flare from the impulsive phase down to below 1~MK. The advantage of their
observations was that they covered the 1097-1119~\AA\ spectral range, which
includes spectral lines formed at all temperatures from 0.01~MK to 15~MK.
The observation lasted enough to allow them to follow the evolution of the
flare plasma from the impulsive phase to well beyond the lower temperature
limit imposed by the use of X-ray instrumentation. They found that the
plasma continued to cool through quasi-equilibrium states until
it reached $\approx$2-3~MK. Below that threshold, Feldman {\em et al.\ } (2003)
reported a dramatic change in the emission, with the surprising, sudden
appearance at the same time of lines from ions formed between 0.01~MK and
1~MK that were not expected to be observed simultaneously in a plasma in
ionization equilibrium. Their observations suggested that the plasma was
undergoing thermal non-equilibrium. Feldman {\em et al.\ } (2003) did not investigate
this final feature of their observation, and limited themselves to reporting
the light curves of the cooling flare plasma. The aim of the present work
is to revisit the observations of Feldman {\em et al.\ } (2003) and investigate in detail
the scenario of thermal non-equilibrium in the final phase of the flare decay,
using a state-of-the-art time-dependent loop hydrodynamic code which includes
non equilibrium ionization. In particular, we focus on predicting the
light curves of all lines emitted in the late phase of flare decay, and on comparing
them to the SUMER observations. We will show that thermal non-equilibrium,
i.e. a catastrophic cooling of a coronal plasma, develops naturally in a standard
flaring loop simulation late in the decay in agreement with observations, and that
it is able to
explain well the sequence and timing of appearance of the observed lines.
This observation represents an excellent testing ground of the
presence of thermal non-equilibrium in the Sun, and it is expected to be
of significant value to study several phenomena other than flares.
In fact, thermal non-equilibrium may produce plasma flows, the formation
of condensations and the loss of connection between electron temperature
and the ionization status of a plasma. Thermal non-equilibrium has been suggested to
be responsible for the formation of prominences from coronal loops (e.g. Antiochos
1980, Karpen {\em et al.\ } 2006 and references therein) and intensity variations traveling
along the loop structure (M\"uller {\em et al.\ } 2005) observed with the \ion[He ii] SOHO/EIT
band (De Groof {\em et al.\ } 2004). Also, it has been recently suggested as a possible
explanation of the main observational properties of coronal loops, including
lifetime, temperature and density profiles, and filamentary structure (Klimchuk
{\em et al.\ } 2010).
In addition, observations of the decay phase of flares have been recently made with
the SDO/EVE instrument (Woods {\em et al.\ } 2010), which observes the emission of the
entire Sun between 1 and 1060~\AA\ at a reduced resolution of 1~\AA\ (60-1060~\AA\
range) and 10~\AA\ (1-60~\AA\ range). EVE flare decays showed qualitatively the
same features described by Feldman {\em et al.\ } (2003), with a series of intensity
enhancements of lines from progressively colder ions, down to around 1~MK.
Since EVE observes continuously the entire solar disk, flare observations are
being made routinely, although at a reduced spectral resolution. Thus, the
high-resolution results of the present work can be used as a guideline for
the interpretation of EVE flare observations.
The observations we use are summarized in Section~\ref{observations}, where
we also refine the measurement of the light curves of all available lines and
improve on the electron density measurements over the results of Feldman {\em et al.\ }
(2003). The hydrodynamic code and numerical setup are described in
Section~\ref{model}, and the comparison between model predictions and
observations is reported in Section~\ref{results}. Section~\ref{discussion}
discusses the results and summarizes this work.
\section{The SUMER picture of a C8 flare}
\label{observations}
\subsection{Observations}
The observations studied in this work are fully described in Feldman
{\em et al.\ } (2003); here we only summarize their main features. The data
set consists of a series of spectra observed by the SUMER spectrometer
on board SOHO (Wilhelm {\em et al.\ } 1995) while its 4\arcsec$\times$300\arcsec\
slit was held fixed with the center pointed at (-980\arcsec,-250\arcsec).
Thus, the entire field of view of the SUMER slit was placed outside
the solar limb, above an active region. The 1097-1119~\AA\ wavelength
range was transmitted to the ground after being observed with an
exposure time of 162.5~s. The choice of such a limited wavelength
range was dictated by the need of a relatively fast cadence, but
it allowed us to include spectral lines from ions formed at all
temperatures of the solar corona. Table~\ref{lines} lists them
along with the temperature of maximum ion abundance from Bryans
{\em et al.\ } (2009); free-free continuum radiation was also observed
during the flare.
The observations were carried out on September 26, 2000 during
the peak of solar cycle 23. They started at 11:41~UT and ended
at 23:07~UT. During the observing run a C8 flare erupted from
the active region and was observed by SUMER. X-ray data from
GOES show that it started at 19:49~UT and by 20:20~UT it had
faded in the X-ray background. Despite the disappearance from
GOES X-ray data, the flare plasma kept evolving: thanks to
the wide range of formation temperature of the lines in the
1097-1119~\AA\ range, SUMER was able to observe the flare for
a much wider time interval, spanning from the very beginning
of the impulsive phase to much beyond the end of the GOES X-ray
flare.
\subsection{Flare emission evolution}
Figure~\ref{time_int} illustrates the behavior of the plasma. Each panel is
built in two steps. First, the intensity of a line is measured at all pixels
along the slit for each of the 162.5~s-long exposures. Then, these 1-D intensity
profiles are placed one next to the other to create a 2-D image, where the
X-axis represents time, and the Y-axis represents the coordinate along the
slit. GOES start and end times of the flare are added to each panel of
Figure~\ref{time_int} as vertical dashed lines. Figure~\ref{time_int}
captures the three main features of the flare:
\begin{enumerate}
\item The impulsive phase shows a sudden, very localized enhancement
of \ion[Si iii] and \ion[O iv] line emission, followed after a few minutes
by a large and long-lasting brightening of \ion[Fe xix] line emission;
\item The decay phase lasts far longer than observed by GOES, indicating
that X-ray instruments only allow to observe a limited portion of the
decay phase of the flare;
\item The final phase of the flare can be observed simultaneously with
\ion[O iv] and \ion[Si iii] (as well as \ion[Ne vi] and \ion[Ca vii],
not shown in the figure).
\end{enumerate}
The presence of two distinct loop sources as responsible of different
phases of a solar flare was remarked by Reale {\em et al.\ } (2004), who modeled
a flare in Proxima Centauri observed with XMM with two different
populations of loops. In a similar way, the
initial \ion[Si iii] and \ion[O iv] pulses observed by SUMER in the 26
Septemper 2000 flare may be due to a different kind of loops than the
rest of the decay phase observed with all the other ions, and thus can
not be modeled together with the latter in a single-loop framework.
Feldman {\em et al.\ } (2003) found that the plasma emitting during the decay
phase of the flare between 20~UT and 21~UT was close to isothermal,
and its temperature decreased exponentially with time from 7.9~MK to
2.5~MK. Then, they reported the simultaneous emission of several ions
whose temperature of formation under ionization equilibrium spanned
more than one order of magnitude. The light curves of these cooler
ions suggest that they are emitted by the cooling post-flare plasma.
In fact, Figure~\ref{time_int} shows that the intensity enhancements
of \ion[Ca x], \ion[O iv] and \ion[Si iii] all overlap, and they prolong
in time those of \ion[Fe xix], \ion[Ca xv] and \ion[Al xi] studied
by Feldman {\em et al.\ } (2003). To provide quantitative measurements of the
relationship among the light curves of all these ions we chose to
concentrate on a single pixel along the SUMER slit, shown in
Figure~\ref{time_int} by the horizontal dashed line. This pixel
corresponds to ``position~38'' in Feldman {\em et al.\ } (2003). We measured
the intensities of all lines in this position as a function of time,
and displayed their normalized values in Figure~\ref{light_curves}.
The impulsive phase of the flare is visible by a sudden brightening
in \ion[Si iii] and \ion[O iv] right before 20~UT, and then the long
decay phase studied by Feldman {\em et al.\ } (2003) is apparent from the peak
of the \ion[Fe xix] line emission to that of \ion[Al xi].
The surprising feature of Figure~\ref{light_curves} is the sudden
appearance (at around 21~UT) and co-existence of the emission from
\ion[Ca x] (formed in equilibrium at $\log T=5.7-6.3$), \ion[Ca vii]
($\log T=5.4-5.9$), \ion[Ne vi] ($\log T=5.4-5.8$), \ion[O iv] ($\log
T=4.9-5.5$) and \ion[Si iii] ($\log T=4.3-5.0$). These ions are emitted
by a previously cooling, near-isothermal plasma undergoing a series of
equilibrium states, and yet cannot be emitted simultaneously
by an isothermal plasma in ionization equilibrium as their temperature
ranges of formation do not overlap. The most likely explanation to the
presence of the emission from these ions is that the cooling post-flare
plasma undergoes a thermal instability as it reaches $\approx 2$~MK.
The modeling effort of the present work is aimed at reproducing the light
curves of the thermal instability after 21~UT. We are particularly interested
at studying the succession of rise and fall of the emission in the colder
ions.
\subsection{Plasma diagnostics}
\label{diagnostics_results}
The plasma temperature was estimated by Feldman {\em et al.\ } (2003) assuming
that the plasma evolved through quasi-equilibrium states from right after
the impulsive phase (around 20:05~UT, approximately 15 minutes after the
GOES beginning of the flare) down to 20:45~UT, so that the shape of the
\ion[Fe xix] and \ion[Ca xv] light curves could be fitted using the
contribution functions of the two lines. The $\log T$ of the plasma
was found to evolve linearly with time from an initial $\log T=6.9$
at 20:05~UT to $\log T=6.40$ at 20:45~UT. The free-free emission was
observed until 21:05~UT and it allowed to determine that the plasma
composition was similar to the standard composition of the solar corona;
this means that the plasma in the decay phase of the flare had a local,
coronal origin rather than being evaporated from the chromosphere.
The plasma density was determined by Feldman {\em et al.\ } (2003) by using the
value of the total emission measure they derived from the free-free
radiation and an assumed value of the total length of the
flaring plasma along the line of sight. However, there are two ions
in the 1097-1119~\AA\ range that emit more than one line: \ion[Si iii]
and \ion[O iv]. The former provides a ratio which is
density sensitive in the $\log N_e=10-12$ range ($N_e$ in cm$^{-3}$),
so that we can attempt a crude estimate using the emissivities from
the CHIANTI database (Dere {\em et al.\ } 1995, 2009). \ion[Si iii] emission
is observed at the very beginning of the flare, during the impulsive
phase, and late in the decay phase. At the beginning of the event
the intensity ratio is 0.60$\pm$0.06, corresponding to $\log N_e > 11.8$
while at the end of
the decay phase the ratio is $0.43\pm0.03$, with little temporal evolution,
indicating a density of $\log N_e =10.5\pm0.3$. The density estimate
made by Feldman {\em et al.\ } (2003) during the impulsive phase
is consistent with the value obtained with the line intensity ratio
only if we assume that the actual volume filled by the plasma is very
small, indicating a filling factor smaller than unity. During the decay
phase, on the contrary, the two estimates agree with each other if we
assume a unity filling factor and a length of the line of sight $H$ in the
1\arcsec-10\arcsec\ range. On Sept. 26, 2000 1\arcsec\ at SOHO's
distance corresponded to $\approx$720~km at the Sun, so that
$7.2\times 10^7 \leq H \leq 7.2\times 10^8$~cm.
\section{Hydrodynamic modeling and numerical setup}
\label{model}
We aim at explaining the evolution of the observed emission through the
hydrodynamic modeling of a flare-heated plasma confined in a coronal
magnetic flux tube as in Reale \& Orlando (2008). We consider
semicircular loops with half-length in the range
$1.5 \times 10^9 < L < 5 \times 10^9$ cm, typical of active region loops,
and constant cross-section area all along the loops. The loops lie on a
plane vertical to the solar surface. It is assumed that the loop plasma
moves and transports energy only along the magnetic field lines. The
plasma evolution can then be described by one-dimensional (1-D)
hydrodynamics (e.g., Peres {\em et al.\ } 1982):
\begin{equation}
\frac{\partial \rho}{\partial t} + \frac{\partial \rho u}{\partial s} = 0
\label{eq1}
\end{equation}
\begin{equation}
\frac{\partial \rho u}{\partial t} + \frac{\partial (P+\rho
u^2)}{\partial s} = \rho g
\end{equation}
\begin{equation}
\frac{\partial \rho E}{\partial t} +\frac{\partial (\rho E+P) u}
{\partial s} = \rho u g - \frac{\partial q}{\partial s} + Q(s, t) -n_e
n_H \Lambda(T)
\end{equation}
\[
\mbox{where \hspace{0.5cm}} E = \epsilon +\frac{1}{2} u^2
\]
\noindent
is the total gas energy (internal energy, $\epsilon$, and kinetic
energy), $t$ the time, $s$ the coordinate along the loop, $\rho =
\mu m_H n_{\rm H}$ the mass density, $\mu = 1.26$ the mean atomic
mass (assuming solar abundances), $m_H$ the mass of the hydrogen
atom, $n_{\rm H}$ the hydrogen number density, $n_{\rm e}$ the
electron number density, $u$ the plasma flow velocity, $P$ the pressure,
$g$ the component of gravity parallel to the field lines, $T$ the
temperature, $q$ the conductive flux, $Q(s,t)$ a function describing
the transient input heating, $\Lambda(T)$ the radiative losses per
unit emission measure (Rosner et al. 1978). The radiative losses
are assumed from plasma in ionization equilibrium and are set
to zero for $T < 20000$ K, to ensure the
detailed energy balance in the chromosphere. Tests
have shown that the simulation results do not change considerably because of
local and temporary changes of the radiative losses function,
as those due by transient deviations from equilibrium of ionization (Reale \&
Orlando 2008). We use the ideal gas law, $P=(\gamma-1) \rho \epsilon$,
where $\gamma=5/3$ is the ratio of specific heats.
The flare evolution may be faster than ionization and recombination
timescales, so that the ionization fraction of the emitting ions may not
have enough time to adjust to the rapidly changing plasma temperature (e.g.
Golub {\em et al.\ } 1989, Orlando {\em et al.\ } 1999).
This effect may be important even in the late phases of the flare.
Therefore in our modeling we also compute the ionization fractions of the
most important elements (namely He, C, N, O, Ne, Mg, Si, S, Ar, Ca, Fe, Ni),
by solving synchronously but independently the continuity equations for each
ion species:
\begin{equation}
\frac{\partial n_i^Z}{\partial t} + \nabla \cdot n_i^Z \mbox{\bf v} =
R_i^Z ~~~~~~~~\begin{array}{l}(Z = 1, ..., N_{elem})\\\\
(i = 1, ..., N_{ion}^Z) \end{array}
\label{eq4}
\end{equation}
\noindent
\[
\mbox{where}~~~R_i^Z = n_e [n_{i+1}^Z\alpha_{i+1}^Z + n_{i-1}^Z
S_{i-1}^Z - n_i^Z(\alpha_i^Z+S_i^Z)]~,
\]
\noindent
$n_{i}^Z$ is the number density of the $i$-th ion of the element $Z$,
$N_{elem}$ the number of elements, $N_{ion}^Z$ the number of ionization
states of element $Z$, $\alpha_i^Z$ the collisional and dielectronic
recombination coefficients, and $S_i^Z$ the collisional ionization
coefficients (Summers 1974).
As discussed in Reale \& Orlando (2008), the heat conduction may become
``flux limited'' during flares (e.g. Brown {\em et al.\ } 1979), and our model
allows for a smooth transition between the classical and saturated
conduction fluxes (Dalton \& Balbus 1993):
\begin{equation}
q = \left(\frac{1}{q_{\rm spi}}+\frac{1}{q_{\rm sat}}\right)^{-1}~,
\end{equation}
\noindent
where $q_{\rm spi}=-\kappa(T)\nabla T$ is the classical conductive
flux (Spitzer 1962), $\kappa(T) = 9.2\times 10^{-7} T^{5/2}$ erg
s$^{-1}$ K$^{-1}$ cm$^{-1}$ the thermal conductivity, $q_{\rm sat}
= -\mbox{sign}\left(\nabla T\right)~ 5\phi \rho c_{\rm s}^3$ the
saturated flux, $c_{\rm s}$ the isothermal sound speed.
The correction factor $\phi=1$ is set according to the values
suggested for the coronal plasma (Giuliani 1984, Borkowski
{\em et al.\ } 1989, Fadeyev {\em et al.\ } 2002, and references therein).
The flare is triggered by a heat pulse defined by the function
$Q(s,t) = H_0 \times g(s) \times f(t)$, where $g(s)$ is a Gaussian function:
\[
g(s) = \exp [ - (s - s_H)^2/ (2 \sigma_H ^2)]
\]
\noindent
and $f(t)$ is a pulse function (see Reale \& Orlando 2008 for details).
In all our simulations, the heat pulse starts at time $t = 0$ and it is
switched off after 180 s. We do not change this duration, since the late
evolution of the flare is governed mostly by the total energy input, which
we tune by changing $H_0$.
The model is implemented and solved numerically using the {\sc flash}\ code
(Fryxell {\em et al.\ } 2000), an adaptive mesh refinement multiphysics code,
extended by additional computational modules to handle the plasma thermal
conduction (see Orlando {\em et al.\ } 2005 for details of the implementation),
the non-equilibrium ionization (NEI) effects (see Reale \& Orlando 2008
for the details), the radiative losses, and the heating function.
As for the initial conditions, before the flare the loop plasma is assumed
to be in pressure and energy equilibrium according to Serio {\em et al.\ } (1981) and
to Orlando {\em et al.\ } (1995). The plasma is very cool and tenuous: the pressure
at the base of the corona is $p_0 = 0.055$ dyn~cm$^{-2}$, which corresponds
to a maximum temperature $T_0 \approx 0.8$~MK at the loop apex, according
to the loop scaling laws (Rosner {\em et al.\ } 1978). Since the pressure increases
enormously during the flare, the flare evolution is largely independent of
these initial conditions. The coronal part of the loop is linked to an
isothermal chromosphere at $T_{\rm c} = 20000$ K and $0.5 \times 10^9$cm
thick, which acts only as a mass reservoir. All the ion species are assumed
to be initially in ionization equilibrium everywhere in the loop.
Since we assume that the loop is symmetric with respect to the vertical
axis across the apex, we simulate only half of the loop. For the
case $L = 3 \times 10^9$ cm, the computational
domain spans a total extension of $3.5 \times 10^9$ cm, including the
chromosphere which has a thickness of $5 \times 10^8$ cm.
At the coarsest resolution, the adaptive mesh algorithm used in the
{\sc flash}\ code ({\sc paramesh}; MacNeice {\em et al.\ } 2000) uniformly covers the
computational domain with a mesh of $16$ blocks, each with $8$ cells.
We allow for 3 levels of refinement, with resolution increasing twice at
each refinement level. The refinement criterion adopted (L\"ohner 1987)
follows the changes in density and temperature. With this grid configuration
the effective resolution is $\approx 3.4\times 10^6$ cm at the
finest level, corresponding to an equivalent uniform mesh of $1024$ grid
points. We use fixed boundary conditions at $s=0$ and reflecting boundary
conditions at $s=s_{\rm max}$ (consistent with the adopted symmetry).
\section{Model results and comparison with observations}
\label{results}
After some exploration of the parameter space, results that provide
a good match to observations are obtained with a loop of half length
$L = 3 \times 10^9$ cm, and with a heat pulse of intensity $H_0 = 8$
erg~cm$^{-3}$ s$^{-1}$, Gaussian width and center $\sigma_H = 10^8$~cm
and $s_H = 8 \times 10^8$ cm, respectively. The heat pulse is therefore
located close the loop footpoint i.e. $\approx 3 \times 10^8$~cm above
the transition region. We now discuss this simulation in some detail.
The evolution of the flare-heated plasma confined in a coronal loop
is well known from previous work (e.g. Nagai 1980, Peres et al. 1982,
Cheng et al. 1983, Nagai \& Emslie 1984, Fisher et al. 1985, MacNeice
1986, Betta et al. 2001). In particular, the choice of the parameters
and the specific model make our results very similar to those described
in Reale \& Orlando (2008) for the case of heat pulse deposited at the
loop footpoints and lasting 180 s. The only difference is that the heating rate
is about 5 times larger, leading to a maximum temperature twice as high,
i.e. about 20 MK. Since here we focus on the late phases of the flare,
we do not enter the details of the initial phases which are similar
to those described by Reale \& Orlando (2008). We only mention that
the loop is rapidly filled by hot plasma and that the plasma begins to
cool down by conduction and radiation as soon as the heat pulse stops,
but the density still grows for a few minutes. The plasma then drains
and cools altogether. Fig.~\ref{fig:evol} illustrates the evolution of
the plasma along half of the model loop from 30 to 50 minutes after the
start of the heat pulse. During this phase, the plasma density at
the formation temperature of the \ion[Si iii] lines is in the range
$10 < \log N_{\rm e} < 11$, in agreement with the values inferred from
spectroscopic analysis of the data (see Sect.~\ref{diagnostics_results}).
The plasma cools down gradually and uniformly along the loop for about 10
minutes but then it suddenly gets thermally unstable and the temperature
drops in about 2 minutes by a factor 10, from $\log T \approx 5.5$
to $\log T \approx 4.5$, in the coronal part of the loop. The density
decreases more gradually and at a constant rate, so that at $t \sim 40$
min the plasma is cold but still relatively dense. Since the plasma is
dramatically out of the hydrostatic equilibrium, it begins to drain more
rapidly and its downward velocity increases from $\sim 20$ km/s to $\sim
100$ km/s. A front of relatively hotter and denser plasma ($\log T \approx 5$)
develops at the loop footpoints and slowly propagates along the loop.
The trigger of the thermal instability can be explained in terms of
the criterion described by Field (1965): below 1~MK the radiative losses
begin to increase for decreasing temperature (negative slope) and the
density is sufficiently high that the cooling becomes catastrophic.
The presence of the instability is even more clear in the evolution of the
loop top temperature. The temperature at the top of the loop is a good
proxy of the loop maximum temperature for most of the time (Fig.~\ref{fig:evol}).
Fig.~\ref{fig:max_tn} shows the loop top temperature and density during
the entire duration of the flare. The heat capacity of the loop is
initially very small and the temperature increases very rapidly, but it
also saturates as well because of the increased losses due to the highly
efficient thermal conduction. After the end of the heat pulse, the
temperature decreases regularly (exponentially) during most of the first
40 min, and then drops suddenly to $T \sim 20000$ K. Below this temperature,
the plasma is assumed to radiate no more, so the thermal evolution for
$t > 41$ min is outside the scope of the modeling. After the temperature
drops, the plasma begins to drain more rapidly and the
density also decreases more rapidly (lower panel).
To study the degree of departure from equilibrium in a plasma, Reale \&
Orlando (2008) introduced for each element an "effective temperature"
(T$_{eff}$), as the temperature for which the equilibrium ion abundances
of the element best approximates the actual values in the plasma. In
case of equilibrium, all elements have the same T$_{eff}$ and this
should equal the electron temperature. Fig.~\ref{fig:max_t} shows
the evolution of T$_{eff}$ at the loop top for the most important ion
species: the ion temperatures substantially differ from $T_{\rm e}$
only during the earliest phase of the flare ($t< 5$ min) and do not
evolve very differently from $T_{\rm e}$ in later phases. We
find again some deviations from the ionization equilibrium around the
thermal instability ($30 < t < 40$ min), because the temperature drops
abruptly. The instability starts from a hotter status ($\sim 1$ MK) for
the light elements C, N, and O ions (when the electron temperature is
$\approx 5\times 10^5$ K). This temperature drop is mainly due to the
fact that the thermal instability occurs in a range of temperature in
which the He-like ionization state of these light elements is the most
populated. Since the temperature $T_{eff}$ is found as the temperature
at which the two most populated ionization states fit those assuming
equilibrium of ionization, the NEI effects together with the fact that
He-like ions are the most abundant in a broad range of temperatures lead
to the observed jump of $T_{eff}$. Note also that $T_{eff}$ for O, N,
C drop consecutively in time because the corresponding He-like ionization
state forms at lower and lower temperatures during the plasma cooling.
Although apparently large, the deviations from equilibrium are so
concentrated in time that they do not produce strong effect on the
visible emission.
The model allows us to predict the light curves of the observed lines.
Fig.~\ref{fig:lc1l} shows them during the development of the
thermal instability in the late phase of the flare, both assuming and not
assuming ionization equilibrium. Time origin and range have been adjusted to allow
a direct comparison with the observed light curves. Fig.~\ref{fig:lc1l}
confirms that the effects of deviations from ionization equilibrium are
not important:
only the light curves of the coldest lines (\ion[Si iii], \ion[O iv]) are
affected, by being slightly broader, i.e. their decay is slightly slower
in the latest phases. Thus, both panels can be compared to observations
and they show the same results: first, the time sequence and the relative
start time of the light curves of each line are both in very good agreement with
observations (within 1-2 min), with the exception of the hottest lines
(\ion[Ca xv]); and second, the observed light curves are considerably
broader than the predicted ones, i.e. both the rise and the decay are
slower than modelled.
We have found that the timing and line sequence could not be reproduced
with the same accuracy with other model parameters, and in particular longer
($5 \times 10^9$ cm) loops, more (to a maximum temperature of 25~MK) or less
(10~MK) intense heating, and also a heat pulse deposited at the loop apex.
We define the root mean square time distance of the simulated and observed
peak between lines as a goodness-of-fit parameter:
\[
\sigma_t = \sum_i \sqrt{ \left( t_{sim} - t_{obs} \right)^2_i}
\]
where $t_{sim}$ and $t_{obs}$ are the times of the peak of the i-th line light
curves measured in the simulation and in the observation, respectively. Here we
align the times to be the same for the peak of the \ion[Si iii] line.
We measure $\sigma_ t$ over the coolest lines (from S xi to Si iii).
We have found that the value of the best-fit model is $\sigma_t \sim 4$
min, while for all others values between 9 and 50 min. For the best-fit
model, each peak of the (cool) model light curves is within 5 min from
respective peak of the observed light curves. Indeed, we have found
$\sigma_t \sim 4$ min also from a simulation with a shorter loop $L = 1.5
\times 10^9$~cm and maximum temperature of 20 MK.
One more constraint from the observation is the absolute intensity of the
line emission. To compute the line intensity from the simulation results we
need to make an assumption on the path length $H$ along the line of sight. For
$L = 3 \times 10^9$ cm, we find that the path length that matches the measured
one is between $1$ and $2 \times 10^8$ cm for the lines marking the thermal
instability (from the warm \ion[S xi] to the cool \ion[Si iii]). This is
in agreement with the estimate of $H$ made in Section~\ref{diagnostics_results},
and it is also compatible with a loop cross-section diameter of the order of
10\% of the loop length, as typical of coronal loops. For the shorter loop
($L = 1.5 \times 10^9$ cm) the path length is larger ($1.5 - 4 \times 10^{8}$ cm)
but still well within the range of values found in Section~\ref{diagnostics_results}.
In this case, the flaring structure would have more the aspect of a flaring loop
arcade. Since we have reasons to believe that the flare occurs in a multi-structured
loops, both scenarios seem possibile. Morevover, these loop lengths are compatible
with the vertical height corresponding to the SUMER slit position above the limb.
Overall, we can say that the loop length, and the heat
intensity and position are relatively well constrained and provide a reasonably
self-consistent scenario.
None of the explored single loop models has been able to reproduce the observed
"broad" light curves, i.e. the observed light curves change more slowly
than those obtained from the hydrodynamic simulations.
The agreement is further improved if we assume that the flaring region is
not a single monolithic loop, but it is substructured into a bundle of
loops or loop strands, each of the same length as the loop we modeled. Each
component is heated at a different time, and the time distribution of the
heating pulses is assumed to be Gaussian with $\sigma=8$~minute width. We
derive the related line light curves simply as the convolution of the single
loop light curves with the (same) Gaussian. The results are shown in
Fig.~\ref{fig:lcok} and compared to observations. The predicted light
curves are now in closer agreement as far as time sequence, time
evolution and time scale are concerned.
We are unable to, and do not pretend to, constrain the number of loops or
strands. The convolution with the Gaussian means that at any time the line
intensities are the same relative to the others (i.e. the ratio is unchanged).
So there is only a scale factor that changes in time, which is a volume factor.
This implies that all loops/strands are heated in the same way, and only the
number of heated loops/strands first progressively increases and then
progressively decreases. We cannot claim that this is the only way to
produce this effect, but we can certainly say that this is a very simple
way to do that, with very few free parameters (only the rate of involved
strands).
Although fine details such as the asymmetry between rise and decay of the
light curves can still be improved, overall we can conclude that this
agreement is evidence of a thermally unstable plasma confined in a
substructured loop system undergoing cooling after a flare.
It is worth noting that the thermal instability occurs at $t \sim 40$
min after the heat pulse, while the observation indicates that it should
occur $\sim 60$ min after the start of the flare (see Fig.2).
The discrepancy between the flare start time and the start of the heat
pulse, as well as the disagreement in the time profile of hot lines, may be
better resolved if the model can produce a temperature time profile that at
first decreases from 20~MK more rapidly but stays longer (by $\simeq$20 min)
above $\simeq$1~MK before going into thermal instability. This appears to be
limited by the current setting of the model for the energy balance.
\section{Discussion and Conclusions}
\label{discussion}
This work aims at explaining the peculiar light curves of several UV lines
in the late phases of a flare observed with SOHO/SUMER. In particular, our
target is the appearance of several spectral lines emitted by plasma
in a wide range of temperature in a relatively short time. The scope of
this work is modeling the post-flare phase and therefore we do not attempt
at reproducing the initial phases of the flare.
We show that light curves are well reproduced by the hydrodynamic evolution
of flaring plasma confined in a closed loop system triggered by a strong
heat pulse. In particular, the sudden, almost simultaneous appearance of
the emission of chromospheric and transition region ions formed more than
one order of magnitude apart in temperature is due to the fact that late
in the cooling phase the plasma becomes thermally unstable, because
the radiative losses increase rapidly below 1 MK, and its temperature
drops rapidly by one order of magnitude in a couple of minutes along most
of the loop. A good relative timing and order of appearance of the light
curves of each line is obtained with a standard flare loop model considering
a loop half-length of $\leq 3 \times 10^9$ cm, and a heat pulse lasting
3~min and deposited in the corona, close to the loop footpoints. The
heating rate is such as to have a flare maximum temperature of about 20~MK.
Better agreement of the duration of the light curve of each line is
obtained if we assume that the flare occurs in a bundle of subloops of
equal length (a stranded loop or a loop arcade). If each substructure is
assumed to be heated at a different time, the predicted light curves can
be lengthened to match the observed ones. This is consistent with the slit
intercepting several concentric arches that undergo independently the same
evolution, so that we see the envelope of this close sequence of events.
The time scale of the distribution of the events is about 8 min, that indicates
a propagation speed of the trigger signal of $\sim 10$ km/s over a length scale
of $\sim 10^9$ cm. This result might be consistent with the presence of fine heat
substructuring during flares (e.g. Antiochos \& Krall 1979, Reeves \& Warren
2002, Warren 2006).
The occurrence of the thermal instability is a necessary ingredient to
explain the observations. However, the effects of such instability on the
ion fractions of the major elements are rather limited, contrarily to our
initial expectations. The reason of such limited departures is the relatively
large density of the cooling plasma, which shortens the ionization and
recombination times scales enough to allow the plasma to efficiently adjust
its ionization status to the rapidly decreasing electron temperature (e.g.
Golub {\em et al.\ } 1989).
We do not reproduce all the details of the observations. In
particular, the light curves of the hot lines are not well reproduced,
and probably would need a more complete analysis of the entire flare
event, which is not the target of this study.
The main limitation of the present work is the lack of spatial information
on the flaring region, and better constraints on the heating location
and timing, and on the properties of the flaring loops systems, can
be obtained by combining the present modeling with flares observed with
the EVE spectrometer and the AIA imaging system on the SDO.
\acknowledgements
The work of Enrico Landi is supported by the NNX10AM17G and NNX11AC20G
NASA grants. Fabio
Reale and Salvatore Orlando acknowledge support from Italian Ministero
dell'Universit\'a e Ricerca and from Agenzia Spaziale Italiana (ASI),
contract I/015/07/0. The software used in this work was in part developed
by the DOE-supported ASC / Alliance Center for Astrophysical Thermonuclear
Flashes at the University of Chicago. The simulations have been executed
at the HPC facility (SCAN) of the INAF-Osservatorio Astronomico di
Palermo. The authors wish to thank the anonymous referee for his/her
comments that helped us improve significantly the paper.
|
\section{Introduction}
A high-efficiency position-sensitive scintillator array for the detection of low energy neutrons has been developed at the National Superconducting Cyclotron Laboratory (NSCL). The Low Energy Neutron Detector Array (LENDA) was built to study charge-exchange (CE) ($p$,$n$) reactions. These reactions have been successfully used in the past to study the spin-isospin response of stable nuclei \cite{Osterfeld92,harakeh01}. LENDA makes possible the extension of such studies to unstable nuclei using the ($p$,$n$) CE reaction in inverse kinematics \cite{Perdikakis09a}. Moreover, LENDA can be utilized in any nuclear reaction study that involves production of low-energy neutrons, for example, for proton-transfer ($d$,$n$) reactions in inverse kinematics and beta-delayed neutron emission experiments. In such investigations, the detection of the slow neutrons with good efficiency as well as energy and angle resolution is essential. LENDA addresses this need by providing fast timing ($\sim$400 ps timing resolution), reasonable position resolution ($\sim$ 6 cm ) and a low neutron detection threshold of 150 keV.\\
For the characterization of the array, a simple test setup utilizing neutrons emitted by a $ ^{252} $Cf radioactive source and photons from a $^{22}$Na source has been used. In this work, the results of the characterization are presented. The setup is described in section \ref{sec:Describe LENDA}. The electronics and data acquisition system (DAQ) is presented in section \ref{sec:DAQ}, a brief description of the data reduction process in section \ref{sec:DataReduc}; the efficiency, timing and position resolution of LENDA in sections \ref{sec:Efficiency}, \ref{sec:Timing} and \ref{sec:Position} respectively. The conclusion follows in section \ref{sec:Conclusion}.
\section{Description of LENDA}
\label{sec:Describe LENDA}
LENDA is an array of 24 neutron detectors.
Each detector module is a type BC-408 \cite{SaG11} plastic scintillator bar with dimensions of $300 \times45 \times 25$ mm.
A Hamamatsu H6410 photomultiplier (PMT) assembly with a photo-electron gain of the order of 10$^{7}$ is used at each end of the bar to detect the scintillation light. In each module, to ensure optimum transmission of light to the photocathode, the PMTs are directly coupled to the scintillator using optical epoxy. The scintillator bar is wrapped with one layer of white nitrocellulose membrane filter paper from Advantec \cite{Adv11}, a layer of aluminum foil and finally black insulating tape to ensure proper light propagation through the bar as well as light-tightness. The Advantec film has comparable performance to other high reflectivity scintillator wrapping materials available commercially (Gore DRP \cite{Gore11}, 3M Vikuity \cite{3MVik11}) and is easier to manipulate while wrapping the LENDA scintillators.\\
\begin{figure}[h]
\centering
\includegraphics[width=0.4\linewidth]{Figure_1.pdf}
\caption{One half of the LENDA array (12 detector modules) mounted on its stand. In this configuration, optimized for charge-exchange (p,n) experiments in inverse kinematics, the bars are arranged at a radius of 1 m from the target position. One stand is placed on each side of the beam line and the array covers roughly 45 degrees of scattering angle in the laboratory frame.}
\label{fig:array}
\end{figure}
Figure \ref{fig:array} shows a picture of one half of the array. Each bar is mounted so that the position-sensitive direction of the bar is vertical, and the shortest side of the bar is parallel to the neutron flight path. At a distance of 1 m from a neutron source, one array covers a solid angle of 0.16 sr in total for a scattering-angle coverage of about 45 degrees.\\
By combining the time and pulse-height information from the PMTs, the timing of a neutron hit, the corresponding scintillation light output, and the neutron's hit position along the longest side of each LENDA bar can be determined. During standard Time-of-Flight (TOF) operation mode, LENDA has to be used with an external detector that provides a time reference signal. In a typical experimental setup where LENDA is used in combination with the S800 spectrometer of NSCL \cite{Baz03}, that signal is generated by charged particles hitting a diamond detector \cite{Stolz06} at the object of the spectrometer.
\section{Data Acquisition system}
\label{sec:DAQ}
A common-stop trigger logic is employed for the time-of-flight measurement. The trigger logic is realized using a Xilinx Virtex-II FPGA chip \cite{Xil11} implemented in a VME-based JTEC XLM72V FPGA module \cite{JTEC11}. Two 32-channel CAEN V792 charge-to-digital converters (QDCs) \cite{Caen11} are used to digitize the charge of the PMT signals. The time information for each event is digitized by two 32-channel CAEN V775 time-to-digital converters (TDCs).
The PMTs are powered by ISEG EHS F030n and 8030n high voltage power supplies. Figure \ref{fig:daq} shows a schematic diagram of the LENDA DAQ. To achieve the highest possible signal-to-noise ratio from the PMTs and thus maintain a low threshold of neutron detection, the PMT voltage is set to be as close as possible to the maximum value of -2700 Volts permitted by the manufacturer. Some voltage variation (50 - 100~Volts in most cases) between individual PMTs is employed to obtain a rough gain-matching of the signals. The anode signal of each PMT is attenuated by a factor of 0.3 using a custom-built 16-channel, two-way splitter-attenuator module, before it is used as an input to the DAQ electronics. This is done to match the signal's pulse-height to the dynamic range of the digitizers and discriminators for the neutron energies of interest.
For each PMT, one of the split anode signals, after being cable-delayed by $\sim 100$ ns, is sent to the QDC to digitize its charge. The other split signal is sent to a Phillips 7106 leading edge discriminator to generate a logic signal that indicates the PMT has fired (PMT-event signal).
The PMT-event signal is sent to the TDC and serves as a start-signal for the TOF digitization of that event.
It is also used as one of the input signals for the FPGA trigger-logic. One more input signal for the logic is required to identify a ``good" event. That signal is provided by an external trigger-detector that is experiment-specific. Five logic signals appear at the output of the FPGA as a result of the trigger processing:
\begin{itemize}
\item A LENDA-hit signal, which signifies that at least one LENDA bar is hit.
This is generated by the coincidence of the PMT-event signals from the two ends of each LENDA bar.
An OR between those coincidences from all LENDA bars gives the LENDA-hit signal.
\item A QDC gate signal provided by the OR of all PMT-event signals.
\item A FAST CLEAR signal generated by the same OR, with a delay of a few hundred ns. This clears the digitizers and aborts the processing of the signal when the DAQ is not triggered by a good event.
\item A VETO signal that eliminates the FAST CLEAR signal whenever a coincidence is realized (DAQ trigger) between the external trigger and the LENDA-hit signal. This allows ``good" data to be digitized successfully.
\item A COMMON STOP signal that causes the TDC to digitize the timing information. This is, in terms of timing, the same as the DAQ trigger signal.
\end{itemize}
\begin{figure}[h]
\includegraphics[width=0.9\linewidth]{Figure_2.pdf}
\caption{A schematic diagram of the DAQ system for LENDA.}
\label{fig:daq}
\end{figure}
The FAST CLEAR scheme that allows events to be selected after the QDC integration has an advantage. The starting time of the QDC integration can be set independently of the timing of the trigger signal. This reduces the amount of delay required to place the detector signal inside the QDC gate.
\section{Data Reduction}
\label{sec:DataReduc}
Neutron TOF, hit position along the longest side of a scintillator bar,
and light output in the scintillator
are obtained by
combining the timing of signals from the PMTs at the top and bottom ends ($t_T$ and $t_B$) of the scintillator and the signal amplitudes ($q_T$ and $q_B$) recorded in the TDC and QDC modules, respectively. Herein, we briefly describe the definition and derivation of these parameters. This analysis is similar as described by other authors~\cite{Pal01, Wak05}.\\
Owing to quenching effects and the propagation time of scintillation light from the hit position in the scintillator to the PMTs,
the signal timing and amplitude are both asymmetric between the top and bottom PMTs. This asymmetry is correlated with the hit position.
Two position-dependent parameters $x_T$ and $x_Q$
are derived from the difference between $t_T$ and $t_B$, and between $q_T$ and $q_B$ using the relations,
\begin{eqnarray}
x_T &\propto& T_{diff}= t_T - t_B, \label{eq:x_T}\\
x_Q &\propto& cogQ = \frac{q_T - q_B}{q_T + q_B}.\label{eq:x_Q}
\end{eqnarray}
The proportionality between $x_{Q}$ and $cogQ$ in equation \ref{eq:x_Q} is a valid approximation in the case of LENDA: The length of the scintillator bars (30~cm) is much smaller than the light attenuation length of 210~cm for the BC-408 material \citep{SaG11}; then, the exponential dependence of $q_T$ and $q_B$ on hit position can be approximated by its Taylor expansion leading to a linear relationship.\\
The neutron TOF ($t$) and light output ($l$), are derived from an average of $t_T$ and $t_B$, and from $q_T$ and $q_B$, respectively, as
\begin{eqnarray}
t&=&\frac{t_T+t_B}{2}-t_{TR},\\
l&=&\sqrt{q_T \times q_B},
\end{eqnarray}
where $t_{TR}$ is the time reference (TR) for TOF provided by the external detector. The $q_T$ and $q_B$ values are converted from QDC channels to light output units of electron equivalent energy (MeV$_{ee}$), by assuming a linear relation between the QDC values and the light output. This approach has been shown to reproduce quite accurately the experimentally measured light output for organic scintillators \cite{Kornilov09}. A typical QDC spectrum for neutrons emitted from a $^{252}$Cf source is shown in figure \ref{fig:QDC_Spectrum}.
\begin{figure}
\centering
\includegraphics[width=0.6\linewidth]{./Figure_3.pdf}
\caption{QDC spectrum of neutrons emitted by a $^{252}$Cf source at a distance of 1m from the detector.}
\label{fig:QDC_Spectrum}
\end{figure}
The relation connecting QDC values to light output is calibrated using the 59~keV $\gamma$-ray of $^{241}$Am and the Compton edge of the 662-keV $\gamma$-ray from a $^{137}$Cs source.\\
\section{Neutron spectrum and efficiency with $^{252}$ Cf}
\label{sec:Efficiency}
Time-of-flight spectra of neutrons from a $^{252}$Cf fission source were measured by LENDA to determine its neutron detection efficiency. The kinetic energy of fission neutrons from $^{252}$Cf ranges approximately from 100~keV to 10~MeV according to the IAEA evaluated neutron energy distribution \cite{Man87a,Man89a}. This distribution was used to reliably determine the energy dependence of the efficiency curve.
\begin{figure}[h]
\centering
\includegraphics[width=0.9\linewidth]{Figure_4.pdf}
\caption{A schematic view of the experimental setup used in the measurement of a $^{252}$Cf neutron spectrum. The $^{252}$Cf fission source was placed at a distance of 1~m from the LENDA bar. Prompt $\gamma$-rays from fission were detected in a liquid scintillator. Neutron-$\gamma$ ray pulse shape discrimination was used to separate them from prompt neutrons.
The $\gamma$-rays identified in the liquid scintillator served as the time reference for the neutron TOF. The shadow bar was used to determine the contribution from background events due to wall-scattered neutrons. See the text for details.}
\label{fig:eff_setup}
\end{figure}\\
In the experimental setup (shown schematically in figure ~\ref{fig:eff_setup}), the $^{252}$Cf source was placed at a distance of 1~m from the LENDA bars. The TR signal for the neutron TOF was established by detecting prompt $\gamma$-rays from fission in a $2^{\prime\prime}\times2^{\prime\prime}$ cylindrical NE-213 type liquid scintillator manufactured by ELJEN (EJ-301) \cite{Elj11a}. This detector was placed at a distance of 6~cm from the source. Pulse shape neutron-$\gamma$ discrimination was employed in the TR scintillator for selecting only $\gamma$-ray events. In all TOF spectra a correction for the ``walk" of the leading edge discriminator was applied in the off-line analysis.
\begin{figure}[h]
\centering
\includegraphics[width=0.6\linewidth]{Figure_5.pdf}
\caption{Response of LENDA to prompt neutrons and $\gamma$-rays from $^{252}$Cf fission events shown as a function of the TOF and light output in the scintillator bar with (a) and without (b) a shadow bar. The projection onto the TOF axis of (a) and (b) is shown in (c).}
\label{fig:PIDToF}
\end{figure}
Figure~\ref{fig:PIDToF}(a) shows a 2-dimensional spectrum of light output versus TOF for $\gamma$ rays and neutrons detected in a LENDA bar. The continuous curve in figure~\ref{fig:PIDToF}(c) shows the projection of the 2-dimensional spectrum on its horizontal (TOF) axis. In figure~\ref{fig:PIDToF}(a), the events included in the sharp peak at $t\sim3$~ns are due to $\gamma-\gamma$ coincidences between the LENDA and TR detectors. The events that lie on the right side of the spectrum at $t > 20$~ns correspond to $\gamma-n$ coincidence events (neutrons detected in LENDA). Events appearing between those two regions are due to $\gamma$ rays from the source, reaching LENDA bars after being scattered from the surrounding walls. Any events in the region of $t < 3$~ns on the left of the $\gamma$ ray peak come from neutrons leaking through the n-$\gamma$ discrimination gate that were being detected in the liquid scintillator in coincidence with $\gamma$ rays in LENDA.\\
Since increasing TOF corresponds to lower neutron kinetic energies, the maximum detected light output in a plastic scintillator like LENDA is varying as a function of TOF. The characteristic light output curve corresponding to full energy deposition by neutron events, becomes practically zero for TOF values of $t \sim 150$~ns. The TOF for this minimum amount of light output corresponds to the smallest kinetic energy that can be detected by LENDA, about 150~keV. It is defined by the lowest threshold setting of the DAQ discriminator module and the maximum gain of the PMTs.\\
The events lying above the light output curve are background events, which are mostly due to neutrons and gammas indirectly reaching LENDA after scattering by the surrounding walls or objects. For these events, TOF is not correlated with the deposited energy. In the offline analysis, the contribution from this background can be reduced by eliminating the events in the region above the light output curve (light-output cut).\\
The remaining background contribution was directly measured by inserting a copper shadow bar to block neutrons directly coming from the source to LENDA as shown in figure~\ref{fig:eff_setup}. The dimensions of the shadow bar were selected so that they match the size of the scintillator bar. It was thick enough (30~cm) to attenuate neutrons with kinetic energy up to 10~MeV, by a factor larger than 600. The effect of the shadow bar blocking part of the background neutron flux was calculated assuming an isotropic background neutron distribution. The shadow bar concealed only 5\% of the total solid angle subtended by LENDA and the systematic error induced in the efficiency calculation was estimated to be of the order of 1\%. The live time of the measurement was used to normalize the background runs to those without the neutron blocker. The spectrum measured with the shadow bar is shown in figure~\ref{fig:PIDToF}(b), where most of the neutron events are due to scattering from walls. The dashed curve in figure~\ref{fig:PIDToF}(c) corresponds to the projection of this 2-dimensional spectrum on the TOF axis.\\
To determine the absolute value of the efficiency, the $^{252}$Cf neutron yield distribution of Refs.~\cite{Man87a,Man89a} was normalized using a yield reference (YR) detector of the same type and size as the TR one. The YR detector was placed at a distance of 40~cm from the source. Its efficiency was simulated using the computer code NEFF7~\cite{Dic98a} and found to vary between 30 and 40\% for neutrons with energies in the region of 1 to 3~MeV. According to previous studies using fission fragment detectors~\cite{cub89}, the results of efficiency simulations using the NEFF7 code are known to be reliable within an error of 5\% for this type of detector.\\
After subtracting the contribution from the background, the number of detected neutrons, $N(E_n)$, as a function of neutron energy, $E_n$, are compared to the neutron yield from the $^{252}$Cf source, $Y(E_n)$ using the relation,
\begin{equation}
\centering
N(E_n) = Y(E_n) \times \epsilon(E_n) \times \Delta\Omega,
\label{eq:eff1}
\end{equation}
where $\Delta\Omega$ is the solid angle of the LENDA bar precisely calculated from the alignment information of the detector and source, $Y(E_n)$ is the standard shape of neutron energy distribution taken from Refs.~\cite{Man87a,Man89a} and normalized using the YR detector, and the $\epsilon(E_n)$ is the neutron-detection efficiency of the LENDA bar as a function of $E_n$.
\begin{figure}[h]
\centering
\includegraphics[width=0.6\linewidth]{Figure_6.pdf}
\caption{Experimental neutron detection efficiencies of a LENDA bar (dots) as a function of neutron energy in comparison with the curve simulated by MCNPX (curves). The error bars (shown where they are larger than the points) indicate statistical uncertainties. See the text for details.}
\label{fig:eff}
\end{figure}
The obtained efficiency curves are shown with dots in Fig.~\ref{fig:eff}, where three threshold levels of 30, 50, and 60 keV$_{ee}$ on the light output were applied in the offline analysis. The results of the simulation of the setup using Monte Carlo N-Particle Transport Code (MCNPX)~\cite{MCNPX02a} are shown for comparison. The thresholds for the recoiling proton energy in the simulations were set to be 200, 300, and 400 keV, respectively, to match the light-output threshold in the experimental data. The simulated distributions reproduce fairly well the trend and magnitude of the measured efficiency for the three values of the threshold. Any small deviations are probably due to ambiguities in background subtraction.
\section{Timing resolution}
\label{sec:Timing}
The timing resolution of LENDA was determined using the two correlated 511 keV photons from the positron annihilation of $ ^{22} $Na. Due to compton-scattering (the dominant mechanism of photon detection in BC-408 at these energies) photons of 511~keV produce a continuum of low-energy events which are equivalent (in terms of pulse height) to the proton recoils produced in the scintillator by 2.3~MeV neutrons. The timing response of the BC-408 plastic scintillator to gammas and neutrons is therefore the same (same pulse shape) with the pulse-height of the compton edge of 511~kev photons corresponding to a neutron energy of 2.3~MeV.\\
\begin{figure}[h!]
\centering
\includegraphics[width=0.5\linewidth]{./Figure_7.pdf}
\caption{Schematic depiction of the setup used to determine the timing resolution of LENDA bars. Photons from a $^{22}$Na positron source were used. In this setup, one of the LENDA bars was functioning as the trigger detector providing the time reference signal for the DAQ. }
\label{fig:setup_resol_22Na}
\end{figure}
For the timing resolution measurements, two LENDA bars were placed opposite to each other so that the source would be at a distance of 1 m from one of the bars, as shown in figure \ref{fig:setup_resol_22Na}. One of the LENDA bars (closest to the source) served as a trigger detector while the other registered the photon TOF.
\begin{figure}[h!]
\centering
\includegraphics[scale=1.0]{./Figure_8.pdf}
\caption{Spectrum representing the timing resolution obtained with LENDA using a $ ^{22}$Na positron emitting source in the arrangement shown in figure \ref{fig:setup_resol_22Na}. The peak corresponds to $\gamma$-$\gamma$ photon coincidence events from positron annihilation that are detected by the two LENDA bars. The width of the peak corresponds to a timing resolution of $\approx$420 ps. }
\label{fig:Resol}
\end{figure}
\begin{figure}[h]
\centering
\includegraphics[width=0.7\linewidth]{./Figure_9.pdf}
\caption{Plot of the energy resolution achievable by LENDA for neutron energies in the range of 0 - 4 MeV, based on measured timing resolution for photons of 511~keV. }
\label{fig:de_resol}
\end{figure}
In the time spectrum of figure \ref{fig:Resol} a prominent peak corresponding to $\gamma$-$\gamma$ coincidences is observed.
The width of the photon peak in figure \ref{fig:Resol} is roughly 600 ps FWHM which corresponds to an intrinsic resolution of 600 ps$ / \sqrt{2} $ $ \approx $ 420 ps for each bar. This timing resolution is nearly constant for light outputs above 200~keV$_{ee}$ and degrades fast for smaller light outputs. As shown in figure \ref{fig:de_resol}, this timing resolution corresponds to an almost constant 5-6\% $ \Delta E/E $ resolution for neutron energies below 4 MeV when detected by LENDA at a distance of 1 m. The degradation of timing resolution for light outputs below 200~keV$_{ee}$ causes to the energy resolution only a small deviation from linearity for low energy neutrons (below roughly 1.3~MeV). In an case this degradation of the timing resolution does not significantly affect the energy resolution of the detector at the corresponding energies.
\section{Position Resolution}
\label{sec:Position}
Using the TDC and QDC information from each PMT as described in section \ref{sec:DataReduc}, the position of an event registered by LENDA can be determined. The PMT closer to the event registers the scintillation light sooner and produces a larger signal than the PMT further away. The position resolution therefore depends on the distance traveled by the scintillation light inside the detector as well as the attenuation caused to the scintillation light while traversing the scintillator material and interacting with the reflective coating of the scintillator. The amount of light produced by an event is also a significant factor and the position resolution is expected to be better for higher deposited energies in the scintillator.\\
The position resolution of LENDA bars was determined experimentally using radioactive sources. In earlier investigations using a prototype detector and a $^{252}$Cf source, a clear correlation was observed between the quantities x$_{T}$ and x$_{Q}$ described in equations \ref{eq:x_T} and \ref{eq:x_Q} \cite{Perdikakis09a}. In figure \ref{fig:position_slideshow}, the 2-dimensional spectra of those two parameters are presented for a source placed at different locations along a detector. The clear shift of the locus of events is directly correlated to the change in location of the source.
\begin{figure}[h]
\centering
\includegraphics[width=0.9\linewidth]{./Figure_10.pdf}
\caption{Events from a source attached on the long side of a LENDA bar generate a locus in the 2-D spectrum of time difference versus charge asymmetry. In this figure the two axes are labeled X$_{T}$ and X$_{Q}$ respectively and calibrated in units of distance (m). The locus of events changes with source location along the scintillator bar, revealing the position sensitivity of the detector. In each figure the distance of the source from the middle point (center) of the bar is displayed as well.}
\label{fig:position_slideshow}
\end{figure}\\
For the position resolution measurements, a bore hole in a large lead block was used to collimate $\gamma$ rays from the $^{22}$Na source. A schematic of the setup is shown in figure \ref{fig:position_setup}. The thickness of the lead ($\approx $10 cm) was enough to block the 511 keV $ \gamma $ rays emitted from $ ^{22} $Na. The diameter of the collimator hole was $1.6$ cm. The collimator was placed at the minimum distance possible from the detector - 5 cm - to minimize the diameter of the irradiation area. The resulting irradiation spot of 2.4 cm total diameter, was small compared to the position resolution expected by such detectors (several cm FWHM). In this setup, LENDA was moved vertically in front of the collimated source for the measurements. The position of the source center along the bar was read from a scale fixed on the side of the detector.\\
\begin{figure}[h]
\centering
\includegraphics[width=0.4\linewidth]{./Figure_11.pdf}
\caption{Picture of the setup used for position calibration}
\label{fig:position_setup}
\end{figure}
For five different locations of the collimated source, T$_{diff}$ and cogQ (defined above in equations \ref{eq:x_T} and \ref{eq:x_Q} respectively) between the 2 PMTs of each bar were recorded. Typical spectra of these quantities are presented in figure \ref{fig:pos_distributions}. The deviation from a Gaussian shape for the distributions in this figure comes from photons scattering on the support structure. In the analysis, these scattering tails where not considered as valid events and a Gaussian fit was used to determine the centroid values of X$_{Q}$ and X$_{T}$ distributions.
\begin{figure}
\centering
\includegraphics[width=0.6\linewidth]{Figure_12.pdf}
\caption{Plot of the T$_{diff}$ (top figure) and cogQ (bottom figure) distributions that correspond to one location of the collimated source along a LENDA bar.}
\label{fig:pos_distributions}
\end{figure}
The results correlating the position of the source with respect to each one of the two parameters are presented in figures \subref*{fig:TdiffvsX} and \subref*{fig:CogvsX} respectively. No constraint was set on the deposited energy of the events and therefore these results correspond to an average position resolution.
\begin{figure}[h]
\centering
\subfloat[Tdiff vs X]{\label{fig:TdiffvsX}\includegraphics[width=0.5\linewidth]{./Figure_13a.pdf}}\hspace{-2em}
\subfloat[cogQ vs X]{\label{fig:CogvsX}\includegraphics[width=0.5\linewidth]{./Figure_13b.pdf}}
\caption{Position dependence of (a) the time difference and (b) of the charge center of gravity between the two PMTs of a LENDA bar. The straight lines in the graphs correspond to a linear fit of the data. The error in the horizontal axis corresponds to the diameter of the aperture used to collimate the source.}
\end{figure}
A linear trend is prominent in both graphs. The small but negligible deviation from linearity observed in the plot of figure \subref*{fig:CogvsX} is attributed to the presence of edge effects that affect the pulse-height of the signal for events very close to the two extremes of the scintillator.
To better understand if the transport of light in the scintillator bars can account for this effect, simulations were performed using the code GUIDE7 \cite{Guide7}. The code was modified to correctly reproduce the transport of light in the area between the scintillator material and its wrapping and to include as a parameter the reflectivity of the wrapping material \cite{Cae08a}.
In figure \ref{fig:guide7_cogq}, the comparison between the simulation results and the experimental data of figure \subref*{fig:CogvsX} are presented. The results of the light transport simulations reproduced fairly well the position dependence observed experimentally.
\begin{figure}[h]
\centering
\includegraphics[width=0.7\linewidth]{./Figure_14.pdf}
\caption{Comparison between simulations of the light transport in LENDA with the code GUIDE7 (red squares) and the experimental data of figure \ref{fig:CogvsX} (black dots). The error in the experimental data due to the aperture size is not included explicitly in this graph but the increased size of the experimental points is used to point out the existence of experimental uncertainty. The dashed straight line is a fit excluding the 2 outermost experimental points, used to guide the eye to the deviation from linearity at the edges of the detector.}
\label{fig:guide7_cogq}
\end{figure}
An average position resolution of 6.3$ \pm $0.2 cm was achieved using the timing information (equation 3) while a slightly worse resolution of 7.2$ \pm $0.3 cm was achieved using the QDC charge signal.
\section{Conclusion}
\label{sec:Conclusion}
LENDA, an array of plastic scintillators capable of detecting low energy neutrons with high efficiency and angular resolution has been built and characterized at NSCL. In this work the results of the characterization of the array using radioactive sources, are presented. The efficiency as well as the timing and position resolutions of LENDA have been determined experimentally. The measurements have been supported by Monte Carlo simulations and general agreement was observed between data and calculations.\\
A low neutron energy threshold of $\sim$150 keV has been demonstrated for LENDA. A timing resolution of $\sim$400 ps and a position resolution of $\sim$6 cm have been obtained. An efficiency greater than 20\% for neutrons below 4 MeV has been measured in the test setup. The results of this work confirm that LENDA is an adequate tool for the study of nuclear reactions in inverse kinematics, especially when low-energy neutron detection is required. Recently, the LENDA array has been used at NSCL to study for the first time the spin-isospin response of the radioactive nucleus $^{56}$Ni \cite{Sas11a}. This was the first experiment with a new measurement technique of (p,n) charge-exchange reactions in inverse kinematics at intermediate energies on unstable isotopes. The results of this investigation proved to be important for improving the description of electron-capture rates on nuclei in the iron region, and for modeling the late evolution of core-collapse and thermonuclear supernovae.
\section*{Acknowledgments}
This work was supported by the US NSF under Grants PHY-0606007 and PHY-0216783 (JINA).
\section*{References}
\bibliographystyle{model1a-num-names}
|
\section{Physics underlying the quantum circuit model}
Computers are physical objects, and computations are physical processes, which may be one of Deutsch's
famous quotes \cite{Deutsch89}. This statement gives rise to a natural question: what is the physics
underlying the quantum circuit model? Deutsch's answer may be that quantum computing
votes for the many-worlds interpretation of quantum mechanics. Nielsen and Chuang's answer to it stays the same
in the two editions of their popular textbook \cite{NC2011}: ``outside the scope of the present discussion".
In our current understanding \cite{Zhang11}, the quantum circuit model can be viewed as a generalization
of the multi-qubit factorisation scattering model in integrable systems \cite{Sutherland04}, and the scalable
quantum computer is thus supposed to be an exactly solvable model satisfying the integrable
condition \cite{Sutherland04}.
It becomes well known in quantum information science after DiVincenzo's work \cite{DiVincenzo95} that
an arbitrary $N$-qubit quantum gate can be expressed as a sequence of products of some two-qubit gates.
Hence a quantum circuit is described as a network of two-qubit gates, which may be called the
{\em locality principle} of the quantum circuit, see Preskill's online lecture notes \cite{Preskill97}.
On the other hand, it has been widely accepted for a long time in integrable systems \cite{Sutherland04}
that many-body factorisable scattering can be expressed as a sequence of two-body scattering. As two-qubit
quantum gates are considered as two-qubit scattering operators, the quantum circuit is thought of
as {\em the generalized factorisable scattering model} defined as a generalization of the factorisation
scattering model \cite{Sutherland04}.
Sutherland in his book \cite{Sutherland04} describes three types of the integrable conditions which are capable
of yielding exactly solvable quantum many-body systems: the first is the Bethe ansatz; the second is the
Yang--Baxter equation; and the third is the quantum Lax equation. In physics, hence, we define {\em integrable
quantum computation} as {\em quantum computing via the integrable condition} so that we can obtain a unified
description on both quantum computing via the Bethe ansatz \cite{Zhang11} and quantum computing via the
Yang--Baxter equation \cite{ZKG05}. Note that we do not deal with quantum computing via the quantum Lax equation
in this paper.
Section 3 is a preliminary introduction on quantum computing via the Bethe ansatz \cite{Zhang11}. The main
reference is Gu and Yang's paper \cite{GY89} on the application of the Bethe ansatz to the $N$ Fermion problem
with delta-functional potential. Given a quantum many-qubit system, i.e., the Hamiltonian is known, if the Bethe
ansatz is satisfied, then it is an integrable model to yield many-qubit factorisable scattering as
a sequence products of two-qubit scattering matrices in which two-qubit gates satisfy the Yang--Baxter equation.
Section 4 is a brief sketch on quantum computing via the Yang--Baxter equation \cite{ZKG05}. Given a nontrivial
unitary solution of the Yang--Baxter equation, the Hamiltonian of an integrable system can be constructed
in principle. Additionally, the paper \cite{ZKG05} is written in the way to emphasize how to derive unitary
solutions of the Yang--Baxter equation from unitary representations of the braid group. In view of the
research \cite{Zhang11}, however, it has been distinct that {\em integrable quantum computation} is a research
subject independent of quantum computing via unitary braid representations. Hence this section is a refinement
of the research \cite{ZKG05} from the viewpoint of {\em integrable quantum computation}.
In mathematics, {\em integrable quantum computation} specifies a type of quantum circuit model of computation in
which two-qubit gates are either the Swap gate (permutation) or nontrivial unitary solutions of the Yang--Baxter
equation.
\section{Quantum computing via the Bethe ansatz}
Let us consider the model of $N$ qubits (spin-1/2 particles) in one dimension interacting via the delta-function
potential, with the Hamiltonian given by
\eq
\label{delta}
H=-\sum_{i=1}^N \frac {\partial^2} {\partial x_i^2} + 2 c \sum_{1\le i <j \le N} \delta(x_i-x_j)
\en
where $c>0$ ($c<0$) means repulsive (attraction) interaction, see \cite{GY89} for the detail. Interested readers
are also invited to refer to Bose and Korepin's article \cite{BK11} for the physical reason why this model is
interesting in experimental quantum information and computation.
For two qubits at $x_1$ and $x_2$, the outgoing wave in the region $x_1 < x_2$
with respective momenta $k_1 < k_2$ is given by
\eq
\label{two-out}
\alpha_{out} \, e^{i (k_1 x_1 + k_2 x_2) }, \quad \alpha_{out}=\alpha_{12},
\en
the incoming wave in the region $x_1 < x_2$ is given by
\eq
\alpha_{21} \, e^{i (k_2 x_1 + k_1 x_2) }, \quad \alpha_{inc}=\alpha_{21},
\en
so that the two-qubit gate (i.e., the scattering matrix) $\check{R}_{12}(k_2,k_1)$ defined by $\alpha_{out}
=\check{R}_{12} \alpha_{inc}$ has the form
\eq
\check{R}_{12}(k_2, k_1) =\frac {i (k_2-k_1)(-P_{12}) +c } {i(k_2-k_1) -c },
\en
where the permutation operator $P_{12}$ exchanges the spins of two flying qubits.
For three qubits at $x_1$, $x_2$ and $x_3$, the outgoing wave in the region $x_1 < x_2 <x_3$
with respective momenta $k_1 < k_2 <k_3$ is given by
\eq
\label{three-out}
\alpha_{out} \, e^{i (k_1 x_1 + k_2 x_2 +k_3 x_3) }, \quad \alpha_{out}=\alpha_{123},
\en
the incoming wave in the region $x_1 < x_2 <x_3$ is given by
\eq
\alpha_{321} \, e^{i (k_3 x_1 + k_2 x_2 + k_1 x_1) }, \quad \alpha_{inc}=\alpha_{321},
\en
so that $\alpha_{out}$ is determined by $\alpha_{inc}$ in the way
\eq
\alpha_{out} = \check{R}_{12}(k_2,k_1)\,\check{R}_{23}(k_3,k_1)\, \check{R}_{12}(k_3,k_2)\,\alpha_{inc},
\en
or in the other way
\eq
\alpha_{out} = \check{R}_{23}(k_3,k_2)\,\check{R}_{23}(k_3,k_1)\, \check{R}_{23}(k_2,k_1)\,\alpha_{inc},
\en
which give rises to the consistency condition (the Yang--Baxter equation) on the three-qubit scattering,
\eq
\label{YBE-three}
\check{R}_{12}(u)\,\check{R}_{23}(u+v)\, \check{R}_{12}(v)\, = \,\check{R}_{23}(v)\, \check{R}_{12}(u+v)\,
\check{R}_{23}(u)
\en
with $u=k_1-k_2$, $v=k_2-k_3$ and $u+v=k_1-k_3$.
The outgoing wave of two qubits at $x_1$ and $x_2$ in the region $x_1 < x_2$ with respective momenta
$k_1 < k_2$ is given by (\ref{two-out}), but the incoming wave in the region $x_2 < x_1$ is given by
\eq
\alpha_{inc} \, e^{i (k_1 x_1 + k_2 x_2) }, \quad \alpha_{inc}=(-P_{12})\,\alpha_{21},
\en
so that the scattering matrix of two qubits in different regions has the form
\eq
\alpha_{out} =\check{R}_{12}(k_2,k_1) (-P_{12})\alpha_{inc}.
\en
Similarly, the outgoing wave of three qubits at $x_1$, $x_2$ and $x_3$ in the region $x_1 < x_2 <x_3$
with respective momenta $k_1 < k_2 <k_3$ is given by (\ref{three-out}), but the incoming wave in the
region $x_3 < x_2 <x_1$ has the form
\eq
\alpha_{321} \, e^{i (k_1 x_1 + k_2 x_2 + k_3 x_3) }, \quad \alpha_{inc}=(-P_{13})\alpha_{321},
\en
where the permutation operator $P_{13}$ exchanges the spins of the first qubit and the third one,
so that the scattering matrix of the three qubits in different regions is given by
\eq
\alpha_{out} = \check{R}_{12}(k_2,k_1)\,\check{R}_{23}(k_3,k_1)\,
\check{R}_{12}(k_3,k_2)\,(-P_{12})\,(-P_{23})\,(-P_{12}) \alpha_{inc}
\en
where $P_{13}=P_{12}\,P_{23}\,P_{12}$ is used.
In term of the notation $\check{R}(u)$, the identity operator $Id$ and the permutation operator $P$,
the two-body qubit gates $\check{R}_{12}$ and $\check{R}_{23}$ have a new form
\eq
\label{ru}
\check{R}(u) = \frac {-1} {i u +c} (c + i u P),\quad \check{R}_{12}=\check{R} \otimes Id,
\qquad \check{R}_{23}=Id \otimes \check{R},
\en
where the two-qubit gate $\check{R}(u)$ satisfies the unitarity condition,
\eq
\check{R}(u)\, \check{R}^\dag(u) = \check{R}^\dag(u)\, \check{R}(u) =Id, \quad \check{R}^\dag(u)=\check{R}(-u).
\en
Hence the multi-qubit factorisable scattering with the delta-function interaction has been
recognized as a quantum circuit model consisting of the Swap gate $P$ and the two-qubit gate
$\check{R}(u)$ (\ref{ru}).
Besides the form (\ref{ru}), the two-qubit gate $\check{R}(u)$ has the other interesting formalism
\eq
\check{R}(\varphi)=-e^{-2 i \varphi} P_- - P_+, \quad P_\pm =\frac {1 \pm P} 2.
\en
with $\varphi=\arctan {\frac u c}$ and $-\frac \pi 2 < \varphi < 0$. At
$u=-c$, the two-qubit gate $\check{R}(u)$ is the $\sqrt{Swap}$ gate given by
\eq
\check{R}(-\frac \pi 4)= - \sqrt{Swap},\quad \sqrt{Swap} = P_{+} + i P_{-}
\en
which is an entangling two-qubit gate but can not yield all
unitary $4 \times 4$ matrices by itself (even with the Swap gate) \cite{LD98}.
Furthermore, the two-qubit gate $\check{R}(\varphi)$ has the form of the time evolutional
operator $U(\varphi)$ of the Heisenberg interaction $\vec{S}_1\cdot \vec{S}_2$ between two
qubits
\eq
\check{R}(\varphi) = -e^{-i \frac \varphi 2} U(-2 \varphi),
\quad U(\varphi)=e^{-i \varphi \vec{S}_1\cdot\vec{S}_2}
\en
modulo a global phase factor with the permutation $P=2 \vec{S}_1\cdot \vec{S}_2 +\frac 1 2$
so that $\check{R}(\varphi)$ and $U(\varphi)$ are two equivalent two-qubit gates in the
quantum circuit model.
As is realized by DiVincenzo et al. \cite{DiVincenzo00}, universal quantum computation
can be set up only using the Heisenberg interaction when a logical qubit is encoded as a
two-dimensional subspace of eight-dimensional Hilbert space of three qubits.
Hence, the quantum circuit model in terms of the Swap gate $P$ and the two-qubit gate
$\check{R}(u)$ (\ref{ru}) is also able to perform universal quantum computation, if
and only if a logical qubit is chosen in a suitable way.
\section{Quantum Computing via the Yang-Baxter equation}
A quantum circuit model in terms of both a nontrivial unitary solution $\check{R}(u)$ of the
Yang--Baxter equation (\ref{YBE-three}) and the Swap gate $P$ is called {\em the generalized
factorisable scattering model} in the paper. Once it is given, the problem becomes how to find out the
underlying physical system, which is thought of as the inverse problem of quantum computing
via the Bethe ansatz.
In the literature, the Yang--Baxter equation has the usual formalism,
\eq
\label{ybe}
(\check{R}(x)\otimes 1\!\! 1_2)\,(1\!\! 1_2\otimes \check{R}(x y))\,(\check{R}(y)\otimes 1\!\! 1_2)
=
(1\!\! 1_2\otimes \check{R}(y))\, (\check{R}(x y)\otimes 1\!\! 1_2)\, (1\!\! 1_2\otimes \check{R}(x))
\en
where the two-qubit gate $\check{R}(x)$ is a linear operator on the Hilbert space of two qubits, i.e.,
$\check{R}: {\cal H} \otimes {\cal H} \to {\cal H}\otimes {\cal H}$, $1\!\! 1_2$ denotes the $2\times 2$
identity matrix, and $x, y$ are called the spectral parameter. Take $x=e^u$ and $y=e^v$, then the formalism of
the Yang--Baxter equation (\ref{YBE-three}) can be derived. As a two-qubit gate, the solution $\check{R}(x)$
of the Yang--Baxter equation (\ref{ybe}) has to satisfy the unitary condition
\eq \label{unitary}
\check{R}(x)\,\check{R}^\dag(x)=\check{R}^\dag(x)\,\check{R}(x)=
\rho 1\!\! 1_4 \en
with the normalization factor $\rho$.
The construction of the Hamiltonian via a nontrivial unitary solution of the Yang--Baxter equation (\ref{ybe})
is a very flexible or subtle process with physical reasoning, namely, there does not exist a universal law to
guide such a construction, partly because the same solution of the Yang--Baxter equation can be yielded
by many different kinds of physical interactions.
The two-qubit gate (\ref{ru}), as a rational solution of the Yang--Baxter equation (\ref{ybe}),
has the form
\eq
\check{R}(\alpha) = - \frac 1 {1+\alpha} (1\!\! 1_4+\alpha P), \quad \alpha= i \frac u c
\en
which is a linear combination of the identity $1\!\! 1$ and the permutation $P$. In terms of the variable $\varphi$,
the two-qubit gate (\ref{ru}) can be expressed as an exponential formalism of the permutation $P$ given by
\eq
\label{rv}
\check{R}(\varphi)=e^{i(\pi-\varphi)} e^{i \varphi P}, \quad \tan\varphi = \frac u c,
\en
hence the Hamiltonian for the generalized factorisable scattering model can be chosen as the Heisenberg
interaction $\vec{S}_1\cdot\vec{S}_2$ between two qubits. On the other hand, it is easy to examine the
delta-function interaction (\ref{delta}) as a suitable Hamiltonian underlying the two-qubit gate (\ref{ru}), but it
is not so explicit to realize the delta-function interaction (\ref{delta}) from the formalism of the two-qubit
gate (\ref{ru}).
In \cite{ZKG05}, an approach for the construction of the Hamiltonian is presented via a two-qubit solution
$\check{R}(x)$ of the Yang--Baxter equation (\ref{ybe}). With an initial state $\psi$, the evolution
state $\psi(x)$ determined by $\check{R}(x)$ is given by
$\psi(x)=\rho^{-\frac 1 2}\check{R}(x)\psi$. The Shr{\" o}dinger equation has the form
\eq i\,\frac {\partial \psi(x)} {\partial x}=H(x)\psi(x) \en with the Hamiltonian $H(x)$ given by
\eq \label{hx} H(x)=i\,\frac {\partial }{\partial x}(\rho^{-\frac 1 2}\check{R})
\rho^{-\frac 1 2}\check{R}^{-1}(x) \en
which is time-dependent of $x$.
In terms of an appropriate parameter instead of $x$, the Hamiltonian $H(x)$ (\ref{hamiltonian}) may be transformed
to a time-independent formalism. For example, the Hamiltonian $H(u)$ (\ref{hamiltonian}) via the two-qubit
gate (\ref{ru}) is obtained to be time-dependent of the parameter $u$ after some algebra using the formula
(\ref{hx}), but the Hamiltonian via the the formalism (\ref{rv}) of this gate is found to be time-independent
Heisenberg interaction.
Consider the other solution of the Yang--Baxter equation (\ref{ybe}),
\eq \label{belltype1}
\check{R}(x)= \left(\begin{array}{cccc}
1+x & 0 & 0 & 1-x \\
0 & 1+x & -(1-x) & 0 \\
0 & 1-x & 1+x & 0 \\
-(1-x) & 0 & 0 & 1+x \end{array}\right), \en
where the unitarity condition (\ref{unitary}) requires $x$ real with the normalization factor $\rho=2(1+x^2)$.
In terms of the variable $\theta$ defined by \eq \label{belltypetran} \cos\theta=\frac 1
{\sqrt{1+x^2}}, \qquad
\sin\theta=\frac x {\sqrt{1+x^2}}, \qquad
\en
the two-qubit gate $\check{R}(x)$ (\ref{belltype1}) has the formalism of $\theta$,
and the Hamiltonian $H(\theta)$ (\ref{hamiltonian}) is calculated to be \eq \label{hamiltonian}
H(\theta)=\frac i 2 \frac {\partial } {\partial \theta}(\rho^{-\frac 1 2}\check
R) \rho^{-\frac 1 2}
\check R^\dag=\frac 1 2\sigma_x\otimes \sigma_y \en
with the conventional form of the Pauli matrices. The time evolutional operator has the form
$U(\theta)=e^{-i H\theta}$ as well as the two-qubit gate (\ref{belltype1}) is given by
\eq \check{R}(\theta)=\cos(\frac \pi 4-\theta)+ 2\,i\,
\sin(\frac \pi 4-\theta)\,H=e^{i (\frac \pi
2-2\,\theta) H}. \en
At $\theta=0$, the two-qubit gate $\check R(0)$ given by
\eq
\check R(0)
= \left( \begin{array}{cccc}
1/\sqrt{2} & 0 & 0 & 1/\sqrt{2}\\
0 & 1/\sqrt{2} & -1/\sqrt{2} & 0\\
0 & 1/\sqrt{2} & 1/\sqrt{2} & 0 \\
-1/\sqrt{2} & 0 & 0 & 1/\sqrt{2}\\
\end{array} \right)
\en
is a unitary basis transformation matrix from the product base to the Bell states.
Besides the $\check{R}(x)$ matrix (\ref{belltype1}), in \cite{ZKG05}, the following type of
nontrivial unitary solutions $\check{R}(x)$ of the Yang--Baxter equation (\ref{ybe}),
\eq \label{matchgate} \check{R}(x) =\left(\begin{array}{cccc}
\omega_1(x) & 0 & 0 & \omega_7(x) \\
0 & \omega_5(x) & \omega_3(x) & 0 \\
0 & \omega_4(x) & \omega_6(x) & 0 \\
\omega_8(x) & 0 & 0 & \omega_2(x)
\end{array}\right),\en
have been recognized as two-qubit gates modulo a phase factor as well as the associated
time-dependent Hamiltonians $H(x)$ (\ref{hx}) also respectively calculated. The two-qubit
gate $\check{R}(x)$ (\ref{matchgate}) is determined by two submatrices $U_1$ and $U_2$,
\eq U_1(x) =\left(\begin{array}{cc}
\omega_1(x) & \omega_7(x) \\
\omega_8(x) & \omega_2(x)
\end{array}\right),\quad U_2(x) =\left(\begin{array}{cc}
\omega_5(x) & \omega_3(x) \\
\omega_4(x) & \omega_6(x)
\end{array}\right), \en
which are in the unitary group $U(2)$.
According to Terhal and DiVincenzo's understanding \cite{TD02} of Valient's work on matchgates, when
the above unitary matrices $U_1$ and $U_2$ are elements of the special unitary group $SU(2)$, quantum
computations with the $\check{R}(x)$ gate (\ref{matchgate}) acting on nearest-neighbor qubits can be
efficiently simulated on a classical computer, whereas the $\check{R}(x)$ gate (\ref{matchgate}) combined
with the Swap gate acting on farther-neighbor qubits may be capable of performing universal quantum
computation. Obviously, the two-qubit gate (\ref{belltype1}) has its two submatrices $U_1$ and $U_2$ in
the $SU(2)$ group.
In terms of $U_1$ and $U_2$ in $SU(2)$, the other two-qubit gate with six nonvanishing entries given by
\eq \check{R}(\theta) =\left(\begin{array}{cccc}
\sinh(\gamma-i\theta)) & 0 & 0 & 0 \\
0 & e^{i\theta} \sinh\gamma & -i \sin\theta & 0 \\
0 & -i\sin\theta & e^{-i\theta}\sinh\gamma & 0 \\
0 & 0 & 0 & \sinh(\gamma+i\theta)
\end{array}\right),\en
with the normalization factor $\rho=\sinh^2\gamma +\sin^2\theta$, can be found in \cite{ZKG05} as a
unitary solution of the Yang--Baxter equation (\ref{YBE-three}) with the spectral parameter $\theta$ and
real parameter $\gamma$, but which gives rise to a time-dependent Hamiltonian (\ref{hamiltonian}).
However, the two-qubit gate (\ref{rv}) associated with the delta-function interaction (\ref{delta}) or
Heisenberg interaction consists of
the following submatrices $U_1$ and $U_2$ given by
\eq U_1 =-\left(\begin{array}{cc}
1 & 0 \\
0 & 1
\end{array}\right),\quad U_2 =-e^{-i\varphi} \left(\begin{array}{cc}
\cos\varphi & i\sin\varphi \\
i\sin\varphi & \cos\varphi
\end{array}\right), \en
where $U_1$ is in $SU(2)$ but $U_2$ is in $U(2)$ because of the global phase. With the encoded logical qubit,
it has been numerically verified \cite{DiVincenzo00} that the two-qubit gate (\ref{rv}) itself can achieve
universal quantum computation by acting on nearest-neighbor qubits.
\section{Concluding remark}
Feynman is well known in the community of quantum information science due to his
pioneering work on both universal quantum simulation in 1982 and quantum circuit model in 1986. Shortly before
he passed away in early 1988, very unexpectedly, Feynman wrote: ``I got really fascinated by these (1+1)-dimensional
models that are solved by the Bethe ansatz", see Batchelor's feature article \cite{Batchelor07} on the research history
of the Bethe ansatz.
A definition of {\em integrable quantum computation} in both physics and mathematics has been proposed in
this article, in order to remove potential conceptual confusions between the papers \cite{Zhang11}
and \cite{ZKG05} as well as declare {\em integrable quantum computation} as an independent research subject in
quantum information and computation. Note that {\em integrable quantum computation} has been also argued from
the viewpoint of the Hamiltonian formalism of quantum error correction codes \cite{Zhang08}.
The asymptotic condition $\check{R}(x=0)$ of the solution of the Yang--Baxter equation (\ref{ybe}) satisfies
the braid group relation and hence $\check{R}(x=0)$ can be viewed as a unitary braiding gate, which suggests
similarities and comparisons between {\em integrable quantum computation} and quantum computing via unitary
braid representations, see \cite{ZKG05, Zhang08} for the detail and \cite{Zhang11} for comments.
\section*{Acknowledgements}
The author wishes to thank Professor Lu Yu and Institute of Physics,
Chinese Academy of Sciences, for their hospitality and support during
the visit in which this work was done and the part of the previous work
\cite{Zhang11} had been done.
|
\section{\label{intro}Introduction}
An effect produced by introduction of structural randomness is perhaps one of the first aspects one would be willing to investigate, once the properties of the model of interest have been successfully studied on regular structures. While computer experiment data keep accumulating for diverse models with structural disorder, this problem is often a real challenge to the theory, though.
We consider the two-dimensional $XY$ model (sometimes referred to as the planar rotator model), which Hamiltonian is traditionally written as
\begin{equation}\label{H_2dXY}
H = - J \sum_{\left<{\bf r,r'}\right>} \cos(\theta_{\bf r} - \theta_{\bf r'})
\end{equation}
with the sum spanning the pairs of nearest neighbors in a square lattice of $N$ sites, $J>0$ being the coupling strength, and the polar angle $\theta_{\bf r}$ representing the only degree of freedom which can be attributed to a spin of unit length rotating in a plane.
The $2D$ $XY$ model is remarkable for its critical properties, as this particular combination of lattice dimensionality and spin symmetry leads to the existence of a finite range of temperatures in which the system exhibits critical-like behaviour [Berezinskii-Kosterlitz-Thouless (BKT) phase] \cite{KosterlitzThouless73, Berezinskii72}; most notably, the spin-spin correlation function decays as a power law with a temperature dependent exponent $\eta=\eta(T/J)$ below the BKT transition point $T_{\mathrm{BKT}}$.
In the low temperature limit, where the spin-wave approximation (SWA) is applicable, i. e. the cosine in the Hamiltonian (1) can be replaced by a quadratic expression without affecting the system properties significantly, one arrives easily at a power law form of the spin-spin correlation function, $R^{-\eta}$, with an exponent linearly dependent on temperature \cite{Rice65,Wegner67}:
\begin{equation}\label{eta_swa}
\eta_{\mathrm{SWA}}=T/2\pi J .
\end{equation}
It is known, however, that as the temperature increases, the real exponent increases non-linearly with temperature, so that it assumes the exact value of $1/4$ at $T_{\mathrm{BKT}}$ \cite{Kosterlitz74}.
Given the two both theoretically and experimentally (computer experiment is meant here) acknowledged facts that the value of the exponent $\eta$ at the BKT transition point cannot be changed by structural dilution (see, for example, \cite{SurunganOkabe05}) whereas the value of the BKT transition temperature is reduced by dilution and depends on its concentration \cite{SurunganOkabe05,KapikranianEtAl08,WysinEtAl05}, one can already make a conclusion that the value of $\eta$ below the BKT point should depend not only on temperature but on dilution concentration as well. It is also clear that $\eta$ should increase with dilution concentration for $T<T_{\mathrm{BKT}}$. It can be interpreted as the increase of effective temperature (decrease of effective interaction) due to dilution..
A number of works have touched this question, mostly using computer simulations. For site dilution, when some fraction of sites is excluded from Hamiltonian (\ref{H_2dXY}), see \cite{WysinEtAl05, BercheEtAl03, KapikranianEtAl07, SunEtAl09}, and for bond dilution, when some fraction of bonds is removed from (\ref{H_2dXY}), see \cite{SurunganOkabe05}.
The present study logically continues the theoretical part of \cite{KapikranianEtAl07}, making a significant advance~\cite{footnote} and covering both the site and bond dilution cases. The focus is on the behavior of the spin-spin correlation function and the searched quantity is the dilution concentration $p$ dependent exponent $\eta$ of the correlation function power-law decay. It is natural to assume that the exponent $\eta=\eta(T/J,p)$ is an analytic function with respect to $p$, away from the percolation threshold. Below, $p$ will denote the fraction of removed bonds or sites, depending on what dilution type is considered. Thus, $\eta$ can be presented as a power series
\begin{equation}\label{eta_series}
\eta(p)\simeq\eta(0)+p(\partial\eta/\partial p)|_{p=0}+\cdots.
\end{equation}
For small dilution concentrations $p$, it is enough to know the slope $(\partial\eta/\partial p)|_{p=0}$ to estimate the value of exponent $\eta$ with good precision. So, in our derivation we drop terms that lead to higher order terms in $p$ in (\ref{eta_series}).
As more simple and transparent from the technical point of view case, bond dilution is considered first in Section \ref{II}, where the spin-spin correlation function is calculated up to the contributions linear in dilution concentration $p$ and temperature $T/J$. The analogous but more technically involved derivation for the correlation function of a system with site dilution can be found in Section \ref{III}. The final results for the exponent of the spin-spin correlation function of the systems with site and bond dilution are given, respectively, by Eqs. (\ref{eta_s.d.}) and (\ref{eta_b.d.}) (see Fig. \ref{fig2}).
\section{\label{II} $2D$ $XY$ model with bond dilution}
In this section the case of bond dilution in the $2D$ $XY$ model is considered.
First, in Subsection \ref{II-1}, the bond diluted spin-wave Hamiltonian and the procedure of configurational averaging are defined. Then, in Subsection \ref{II-2}, the spin-spin correlation function is calculated up to the contributions linear in dilution concentration $p$ and temperature.
\subsection{\label{II-1} Bond diluted Hamiltonian and configurational averaging}\label{B.d. Hamilt. and conf.avrg.}
Hamiltonian (\ref{H_2dXY}) in the SWA and with bond dilution can be written as
\begin{equation}\label{H_b.d.}
H_{\mathrm{b.d.}} = \frac{J}{2} \sum_{\bf r}\sum_{\alpha=x,y} (\theta_{\bf r} - \theta_{{\bf r}+{\bf u}_\alpha})^2 (1 - p_{{\bf r},\alpha}),
\end{equation}
where ${\bf u}_x = (a,0)$, ${\bf u}_y = (0,a)$ ($a$ is the lattice spacing), and $p_{{\bf r},\alpha}=1$ if bond $({\bf r,u}_\alpha)$ is removed and $0$ otherwise (see Fig. \ref{fig1}). Then, any thermodynamic quantity characterizing the system will depend on the particular choice of configuration $\{p_{{\bf r},\alpha}\}$ of the discrete variables.
One is willing to consider here what is often referred to as quenched dilution, i. e. when there is a fixed fraction $p$ of removed bonds distributed randomly in the system and frozen at their position \cite{Brout59}. Meaningful physical quantities can be obtained averaging them over the configurations with a fixed fraction of removed bonds $p$. For a large system one might as well allow \emph{all} configurations, ascribing them a probabilistic weight
\begin{eqnarray}
P(\{p_{{\bf r},\alpha}\}) = \prod_{{\bf r},\alpha} \left[(1-p)(1-p_{{\bf r},\alpha}) + p p_{{\bf r},\alpha}\right]\nonumber\\ = (1-p)^{\sum_{{\bf r},\alpha}(1-p_{{\bf r},\alpha})}p^{\sum_{{\bf r},\alpha}p_{{\bf r},\alpha}},
\end{eqnarray}
meaning that a bond is removed with probability $p$, which will lead to the fact that only realizations with fraction $\sum_{{\bf r},\alpha}p_{{\bf r},\alpha}/(2N) \simeq p$ ($2N$ is the number of bonds in the initial lattice) of removed bonds will make essential contribution to the averaged quantities, when $N\to\infty$. It immediately follows that
\begin{equation}\label{p^i_avrg}
\overline{p^i_{{\bf r},\alpha}} = p,\quad \overline{p_{{\bf r}_1,\alpha_1} \cdots p_{{\bf r}_i,\alpha_i}} = p^i
\end{equation}
(all pairs $({\bf r}_1,\alpha_1), \ldots, ({\bf r}_i,\alpha_i)$ are different), where $\overline{(\ldots)}$ means averaging with respect to disorder configurations,
$$
\overline{(\ldots)} = \left(\prod_{{\bf r},\alpha}\sum_{p_{{\bf r},\alpha}=0,1}\right) P(\{p_{{\bf r},\alpha}\}) \ldots ,
$$
hereafter referred to as configurational averaging.
\begin{figure}[h]
\center{\includegraphics[width=0.35\textwidth,angle=0]{fig1.eps}}
\caption{\label{fig1} The occupation number $p_{{\bf r},\alpha}$ ($\alpha = x,y$) takes value $1$ if bond $({\bf r},{\bf u}_\alpha)$ is removed and $0$ otherwise.}
\end{figure}
It is convenient to rewrite Hamiltonian (\ref{H_b.d.}) in the Fourier transformed variables $\theta_{\bf k} = \frac{1}{\sqrt{N}}\sum_{\bf r}e^{i{\bf kr}}\theta_{\bf r}$ as
\begin{equation}\label{H_b.d._four}
H_{\mathrm{b.d.}} = H_0 + H(\{p_{{\bf r},\alpha}\}), \ H(\{p_{{\bf r},\alpha}\})\equiv\sum_{{\bf r},\alpha}p_{{\bf r},\alpha} H_\alpha({\bf r}),
\end{equation}
where
\begin{equation}\label{H_pure_four}
H_0 = -J\sum_{{\bf k}}\gamma_{\bf k}\theta_{\bf k}\theta_{\bf -k}
\end{equation}
with
\begin{equation}\label{gamma}
\gamma_{\bf k} = 2\left(\sin^2\frac{k_xa}{2} + \sin^2\frac{k_xa}{2}\right)
\end{equation}
is the Hamiltonian of the undiluted system, and
\begin{equation}\label{H_alpha(r)}
H_\alpha({\bf r}) = -\frac{J}{2N}\left[\sum_{{\bf k}}e^{-i{\bf kr}}(1-e^{-ik_\alpha a})\theta_{\bf k}\right]^2 .
\end{equation}
The sums over ${\bf k}$ in (\ref{H_pure_four}) and (\ref{H_alpha(r)}) span the 1st Brillouin zone.
The thermodynamic average of some physical quantity $A$ can be written as
\begin{equation}\label{<A>}
\left<A\right> = \textrm{Tr}_\theta Ae^{-\beta H_{\mathrm{b.d.}}}/\textrm{Tr}_\theta e^{-\beta H_{\mathrm{b.d.}}}
\end{equation}
Since $\theta_{\bf k}$ is a complex variable (for ${\bf k} \neq 0$): $\theta_{\bf k} = \theta^c_{\bf k} + i \theta^s_{\bf k}$, $\mathrm{Tr}_{\theta}$ above means
\begin{equation}\label{Tr_phi}
\mathrm{Tr}_{\theta} = \int d\theta_0 \prod_{{\bf k}\in
B/2}\int_{-\infty}^\infty d\theta^c_{\bf k}
\int_{-\infty}^\infty d\theta^s_{\bf k}\ ,
\end{equation}
where $B/2$ stands for a half of the 1st Brillouin zone excluding ${\bf k} = 0$ ($\theta^c_{\bf k}$ and $\theta^s_{\bf k}$ in the other half are not independent due to the relations: $\theta^c_{\bf -k} = \theta^c_{\bf k}$ and $\theta^s_{\bf -k} = -\theta^s_{\bf k}$). Note, that it is possible to extend the bounds of integration in (\ref{Tr_phi}) to infinity, since the functions that stand after the trace are always rapidly decaying at $\beta J\to\infty$.
The configurationally averaged value of $\left<A\right>$ can be obtained using the Taylor series representations of the exponential and $(1+x)^{-1}$ functions with respect to powers of $H(\{p_{{\bf r},\alpha}\})$. The equalities in (\ref{p^i_avrg}) easily lead to
\begin{widetext}
\begin{eqnarray}\nonumber
&&\overline{H^i(\{p_{{\bf r},\alpha}\})} = p\sum_{{\bf r},\alpha} H^i_\alpha({\bf r})
+p^2 \Bigg[\sum_{{\bf r},\alpha} \sum_{{\bf r}',\alpha'}\Bigg]' \frac{i!}{2}\sum_{i'=1}^{i-1} \frac{H_\alpha^{i-i'}({\bf r}) H_{\alpha'}^{i'}({\bf r}')}{(i-i')!i'!} + \cdots \\\label{expansion_in_p} &&
+p^n \Bigg[\sum_{{\bf r}_1,\alpha_1}\cdots \sum_{{\bf r}_n,\alpha_n}\Bigg]'\frac{i!}{n!}\sum_{i_1=1}^{i-1}\sum_{i_2=1}^{i_1-1} \cdots \sum_{i_{n-1}=1}^{i_{n-2}-1}
\frac{H_{\alpha_1}^{i-i_1}({\bf r}_1) H_{\alpha_2}^{i_1-i_2}({\bf r}_2)\cdots H_{\alpha_{n-1}}^{i_{n-2}-i_{n-1}}({\bf r}_{n-1}) H_{\alpha_n}^{i_{n-1}}({\bf r}_n)}{(i-i_1)!(i_1-i_2)!\cdots (i_{n-2}-i_{n-1})!i_{n-1}!} + \cdots ,
\end{eqnarray}
\end{widetext}
where $[...]'$ means that the terms having any coinciding pairs of indexes, ${\bf r}_i = {\bf r}_j, \alpha_i = \alpha_j$, are excluded from the sums enclosed in brackets. This result will be applied in the next subsection to calculate the spin-spin correlation function.
\subsection{\label{II-2} Spin-spin correlation function of the bond diluted $2D$ $XY$ model}
The spin-spin correlation function of the $XY$ model described by Hamiltonian $H$ can be written as
\begin{equation}\label{G(R)_def}
G({\bf R}) = \Re \left<e^{i(\theta_{\bf R} - \theta_0)}\right> =
\Re \frac{\mathrm{Tr}_\theta e^{-\beta H + i\sum_{\bf k}\eta_{\bf
k}({\bf R})\theta_{\bf k}}}{\mathrm{Tr}_\theta e^{-\beta H}}
\end{equation}
with
\begin{eqnarray}\label{eta_def}
\eta_{\bf k} ({\bf R}) &=& \left( e^{-i\bf kR} - 1
\right)/\sqrt{N} .
\end{eqnarray}
For the undiluted system, Eq. (\ref{H_pure_four}), one can write, since $\theta^c_{\bf -k} = \theta^c_{\bf k}$ and $\theta^s_{\bf -k} = -\theta^s_{\bf k}$, using the notations of (\ref{Tr_phi}),
\begin{eqnarray}\nonumber
&&G_0({\bf R}) = \Re \mathrm{Tr}_\theta e^{-2\beta J\sum_{{\bf k}\in
B/2}\gamma_{\bf k}\left[(\theta^c_{\bf k})^2 + (\theta^s_{\bf k})^2\right]}
\\\nonumber
&&\times e^{2i\sum_{{\bf k}\in B/2}(\eta^c_{\bf k}\theta^c_{\bf k} - \eta^s_{\bf k}\theta^s_{\bf k})} / \mathrm{Tr}_\theta e^{-2\beta J\sum_{{\bf k}\in
B/2}\gamma_{\bf k}\left[(\theta^c_{\bf k})^2 + (\theta^s_{\bf k})^2\right]},
\\\label{G(R)_four}
\end{eqnarray}
where $\eta^c_{\bf k}$ and $\eta^s_{\bf k}$ denote the real and imaginary parts of $\eta_{\bf k} ({\bf R})$. It is straightforward to get from the Gaussian integration:
\begin{equation}\label{G0(R)}
G_0({\bf R}) = \exp\left[-\frac{1}{4\beta J}\sum_{{\bf k}\neq
0} \eta_{\bf k} ({\bf R}) \eta_{\bf -k} ({\bf R})/ \gamma_{\bf k}\right],
\end{equation}
here and below sums over $\bf k$ span the entire 1st Brillouin zone except the point ${\bf k} = 0$.
To obtain the asymptotic behaviour of (\ref{G0(R)}) at $R\to\infty$ one should use the fact that $\eta_{\bf k} \eta_{\bf -k} = \frac{4}{N}\sin^2\frac{\bf kR}{2}$ oscillates very fast comparing to $1/\gamma_{\bf k}$ and, thus, can be replaced by its average value $2/N$ everywhere expect the region close to the singularity point ${\bf k}=0$. In this region, replacing in the thermodynamic limit $N\to\infty$ the sum with an integral and taking the leading terms of the Taylor expansion of $\sin^2\frac{\bf kR}{2}$ and $\gamma_{\bf k}$, one gets an integrable expression. One arrives at (see, for example, \cite{Wegner67} or \cite{KapikranianEtAl07} for details)
\begin{equation}\label{sum_etaeta/gamma}
\sum_{{\bf k}\neq
0} \eta_{\bf k} ({\bf R}) \eta_{\bf -k} ({\bf R})/ \gamma_{\bf k} \underset{R\to\infty}{\rightarrow} \frac{2}{\pi}\ln\frac{R}{a} + \mathrm{const}.
\end{equation}
It is easy to see that this asymptotic expression leads to a power-law decay of the spin-spin correlation function, $R^{-\eta}$, with an exponent given by (\ref{eta_swa}).
For a system with bond dilution the spin-spin correlation function is given by (\ref{G(R)_def}) with $H=H_{\mathrm{b.d.}}$, Eqs. (\ref{H_b.d._four})-(\ref{H_alpha(r)}). Applying the scheme of configurational averaging described in Subsection \ref{B.d. Hamilt. and conf.avrg.} to the correlation function, one is able to collect the resulting series into the following expression:
\begin{eqnarray}\nonumber
\overline{G({\bf R})} &=& G_0({\bf R})\Bigg\{1 + p \sum_{{\bf r},\alpha}
\Bigg(\left<e^{i\sum_{\bf k} \eta_{\bf k} \theta_{\bf
k}} e^{-\beta H_\alpha ({\bf r})}\right>_0
\\\label{overline_G_b.d.}
&&\times G^{-1}_0({\bf R})\left<e^{-\beta H_\alpha ({\bf r})}\right>^{-1}_0 - 1 \Bigg) + O(p^2)\Bigg\},\quad
\end{eqnarray}
where the terms of higher order in $p$ are dropped and $\left<\ldots\right>_0$ denotes thermodynamic averaging with Hamiltonian (\ref{H_pure_four}) of the undiluted system:
\begin{equation}\label{<...>_0}
\left<\ldots\right>_0 = \textrm{Tr}_\theta e^{-\beta H_0}\ldots / \textrm{Tr}_\theta e^{-\beta H_0}.
\end{equation}
Now, using the Taylor series representation of an exponential and the results of Appendix \ref{AppendxA} [Eqs. (\ref{theta...theta_0}) and (\ref{empty_sum_1})], one obtains for $H_\alpha ({\bf r})$ given by (\ref{H_alpha(r)})
\begin{equation}\label{avrg1}
\left<e^{-\beta H_\alpha({\bf r})}\right>_0 = 1 + \sum_{n=1}^{\infty} \frac{(2n-1)!!}{(2n)!!} \left(\frac{1}{2}\right)^n = \sqrt{2}
\end{equation}
(here and below $(2m)!!\equiv \prod_{i=1}^{m} 2i$, $(2m-1)!!\equiv \prod_{i=1}^{m} (2i-1)$, $m=1,2,\ldots$, and $0!!\equiv1$).
In a similar way, using (\ref{theta...theta_e^eta_0}), (\ref{empty_sum_1}), and the notation
\begin{equation}
I_\alpha({\bf r}) \equiv \frac{1}{\sqrt{N}} \sum_{\bf k} e^{-i{\bf kr}}\left(1 - e^{-ik_\alpha a}\right)\eta_{-\bf k}/\gamma_{\bf k} ,
\end{equation}
one arrives at
\begin{eqnarray}\nonumber
\left<e^{i\sum_{\bf k} \eta_{\bf k} \theta_{\bf k}}e^{-\beta H_\alpha({\bf r})}\right>_0 = G_0({\bf R})\Bigg\{ 1 + \sum_{n=1}^{\infty}\sum_{l=0}^n \frac{(-1)^{n-l}}{(2\beta J)^{n-l}}
\\\label{avrg2a}
\times \frac{(2n-1)!!}{(2l)!!(2n-2l)!} \left(\frac{1}{2}\right)^n I_\alpha^{2(n-l)}({\bf r})\Bigg\} .\qquad
\end{eqnarray}
The unity and the term with $l=n$ in (\ref{avrg2a}) give $\sqrt{2}$ [see (\ref{avrg1})]. Changing index $n\to i=n-l$ and rearranging the terms of the infinite series, one has
\begin{eqnarray}\nonumber
&&\left<e^{i\sum_{\bf k} \eta_{\bf k} \theta_{\bf k}}e^{-\beta H_\alpha({\bf r})}\right>_0 = G_0({\bf R})\Bigg\{ \sqrt{2} + \sum_{i=1}^{\infty} \frac{(-1)^{i}}{(4\beta J)^{i}}
\\\label{avrg2b}
&&\times \frac{I^{2i}_\alpha({\bf r})}{(2i)!} \sum_{l=0}^{\infty} \frac{(2(l+i)-1)!!}{(2l)!! 2^l} \Bigg\} = G_0({\bf R})\sqrt{2} e^{- \frac{I_\alpha^2({\bf r})}{4\beta J}} .\quad
\end{eqnarray}
The Taylor series representation of $(1 - x)^{-n/2}$,
\begin{equation}
(1 - x)^{-n/2} = 1 + \sum_{l=1}^{\infty} \frac{(2l-2+n)!!}{(2l)!!} \frac{n}{n!!} x^l,
\end{equation}
with $x=1/2$ and $n=1$ and $2i+1$ was used in (\ref{avrg1}) and (\ref{avrg2b}), respectively.
Now, having (\ref{avrg1}) and (\ref{avrg2b}), one can write the spin-spin correlation function in the low temperature limit as
\begin{eqnarray}\nonumber
\overline{G({\bf R})} &=& G_0({\bf R})\Bigg\{1 - \frac{p}{4\beta J} \sum_{{\bf r},\alpha}I_\alpha^2({\bf r}) \Bigg\}
\\\label{b.d.corr.funct.}
&\simeq & G_0({\bf R}) e^{- \frac{p}{4\beta J} \sum_{{\bf r},\alpha}I_\alpha^2({\bf r})}.
\end{eqnarray}
Noticing that $\sum_{{\bf r},\alpha}I_\alpha^2({\bf r}) = 2\sum_{\bf k} \eta_{\bf k} \eta_{-\bf k}/\gamma_{\bf k}$, from (\ref{sum_etaeta/gamma}) immediately follows a power law decay of the correlation function, $R^{-\eta}$, with a dilution concentration dependent exponent
\begin{eqnarray}\nonumber
\eta_{\mathrm{b.d.}}(p) &=& \eta(0) +p\frac{T}{\pi J} + O(p^2) + O((T/J)^2)
\\\label{eta_b.d.}
&\simeq& \eta(0)\left(1 + 2 p\right),
\end{eqnarray}
where $\eta(0)$ is the exponent of the pure system, Eq. (\ref{eta_swa}).
\section{\label{III} $2D$ $XY$ model with site dilution}
In this section the case of site dilution in the $2D$ $XY$ model is considered.
In Subsection \ref{III-1}, the site diluted spin-wave Hamiltonian is defined, then, in Subsection \ref{III-2}, the spin-spin correlation function is calculated up to the contributions linear in dilution concentration $p$ and temperature.
\subsection{\label{III-1}Hamiltonian of the $2D$ $XY$ model with spin
vacancies}
The spin-wave Hamiltonian of a system with site dilution differs from that of bond dilution, Eq. (\ref{H_b.d.}), in the way that the {\it four} bonds adjacent to each spinless site must be removed, so the occupation number
\begin{equation}
p_{\bf r} = \Big\{\begin{array}{ll}1,\ \text{if there is no spin on site}\ {\bf r};\\0,\ \text{otherwise,}\end{array}
\end{equation}
has to be introduced; then,
\begin{equation}\label{H0+H1}
H_{\text{s.d.}} = H_0 + H(\{p_{\bf r}\}), \quad H(\{p_{\bf r}\}) = \sum_{\bf r} p_{\bf r} H_1({\bf r}),
\end{equation}
where $H_0$ is the Hamiltonian of the pure model, Eq. (\ref{H_pure_four}), and
\begin{equation}\label{H_1(r)}
H_1({\bf r}) = - \frac{J}{2} \sum_{\bf u} (\theta_{\bf r} -
\theta_{\bf r+u})^2
\end{equation}
with ${\bf u} = (\pm a,0), (0,\pm a)$, which in the Fourier variables reads as
\begin{equation}\label{H_1(r)_fourier}
H_1({\bf r}) = \frac{J}{N}\sum_{\bf k,k'} e^{-i{\bf (k+k')r}} g_{\bf k,k'} \theta_{\bf k} \theta_{\bf k'}
\end{equation}
with $g_{\bf k,k'} = \gamma_{\bf k,k'} - \gamma_{\bf k} - \gamma_{\bf k'}$ [$\gamma_{\bf k}$ was defined in (\ref{gamma})].
One can notice that expression (\ref{H0+H1}) is not precise when there are neighboring spin vacancies; in this case, the common bond between the vacant sites is subtracted from the ``pure" Hamiltonian twice, so it is, in fact, brought back with an opposite sign. The precise form of $H(\{p_{\bf r}\})$ would be
\begin{equation}\label{H0+H1+H2}
H(\{p_{\bf r}\}) = \sum_{\bf r} p_{\bf r} H_1({\bf r}) + \sum_{\left<{\bf r,r}'\right>}p_{\bf r}p_{\bf r'}H_2({\bf r,r}'),
\end{equation}
where $H_2({\bf r,r'}) = \frac{J}{2} (\theta_{\bf r} - \theta_{\bf r'})^2$. However, it is not only that the second term in (\ref{H0+H1+H2}) gives contributions of order of $p^2$ and higher, after configurational averaging, but it can be {\it always} dropped when considering the spin-spin correlation function, since any non-physical extra bonds corresponding to neighboring spinless sites in (\ref{H0+H1}) are isolated from the rest of the system.
\subsection{\label{III-2} Spin-spin correlation function of the site diluted $2D$ $XY$ model}
Now, everything said in Section \ref{B.d. Hamilt. and conf.avrg.} about the bond dilution and configurational averaging can be applied to site dilution as well with the only difference that here occupation numbers $p_{\bf r}$ are defined for each site ${\bf r}$, and $p = \overline{p_{\bf r}} \simeq \sum_{\bf r}p_{\bf r}/N$ is now the fraction (concentration) of removed {\it sites}.
Then, dropping the higher order terms with respect to dilution concentration $p$, the configurationally averaged correlation function can be written as
\begin{eqnarray}\nonumber
\overline{G({\bf R})} = G_0({\bf R})\Bigg\{1 + p \sum_{\bf r}
\Bigg(\left<e^{i\sum_{\bf k} \eta_{\bf k} \theta_{\bf
k}} e^{-\beta H_1({\bf r})}\right>_0
\\\label{G2(R)_avrg}
\times G^{-1}_0({\bf R})\left<e^{-\beta H_1({\bf r})}\right>^{-1}_0 - 1 \Bigg) +O(p^2)\Bigg\} \
\end{eqnarray}
with $\eta_{\bf k}$ given by (\ref{eta_def}).
The thermodynamic averages in (\ref{G2(R)_avrg}) can be calculated
using the Taylor series expansion: $e^{-\beta H_1({\bf r})} = \sum_{n=0}^{\infty} (-\beta H_1({\bf r}))^n/n!$. Then, the problem reduces to the calculation of the quantity $\left< e^{i\sum_{\bf k} \eta_{\bf k} \theta_{\bf k}} H^n_1({\bf r}) \right>_0$ with $\eta_{\bf k}$ given by (\ref{eta_def}) and $\eta_{\bf k} = 0$, which is presented in Appendix \ref{AppendxB}. Looking at the results (\ref{exp_1}) and (\ref{exp_2}), it is easy to see that
\begin{eqnarray}\nonumber
\left<e^{-\beta H_1({\bf r})}\right>_0 = \prod_{i=1}^{\infty}\sum_{l=0}^{\infty} \frac{1}{l!} \left((-1)^i\frac{I_i}{2i}\right)^l
= \exp\sum_{i=1}^{\infty}(-1)^i \frac{I_i}{2i},
\end{eqnarray}
and, similarly:
\begin{eqnarray}\nonumber
\left<e^{i\sum_{\bf k} \eta_{\bf k} \theta_{\bf k}}e^{-\beta H_1({\bf r})}\right>_0 &=& G_0({\bf R}) \exp \sum_{i=1}^{\infty}(-1)^i \frac{
I_i}{2i}
\\\nonumber
&&\times \exp\left[{-\frac{1}{4\beta J}\sum_{j=1}^{\infty}(-1)^j I^*_j}\right].
\end{eqnarray}
Explicit expressions for the quantities $I_i$ and $I^*_i$ are given in (\ref{I_i_def})-(\ref{tilde_I_i}).
Finally, from (\ref{G2(R)_avrg}),
\begin{equation}\label{G2(R)_2}
\overline{G({\bf R})} = G_0({\bf R})\left\{1 + p \sum_{\bf r}
\left(e^{-{\frac{1}{4\beta J}\sum_{j=1}^{\infty}(-1)^j I^*_j}} - 1
\right)\right\}.
\end{equation}
Using the result of Appendix \ref{AppendxC}, Eq. (\ref{sum_I*_i}), with $\eta_{\bf k}$ given by (\ref{eta_def}), one has
\begin{eqnarray}\nonumber
\overline{G({\bf R})} = G_0({\bf R})\Bigg\{1 -2p + p \sum_{{\bf
r}\neq 0,{\bf R}} \Big(e^{- \frac{\pi}{8\beta J}
F_1({\bf r},{\bf R})}
\\\label{G2(R)_3}
\times e^{- \frac{\pi}{8\beta J(\pi-2)} F_2({\bf r},{\bf R})}
- 1 \Big)\Bigg\},
\end{eqnarray}
where
\begin{eqnarray}\nonumber
F_i({\bf r},{\bf R}) &=& \left[S_i(x-X,y-Y) - S_i(x,y)\right]^2
\\\label{F_i}
&&+ \left[S_i(y-Y,x-X) -S_i(y,x)\right]^2
\end{eqnarray}
($ i=1,2$) with the functions $S_1$, $S_2$ defined in (\ref{S1_def}),
(\ref{S2_def}).
Now, one can expand the exponential function, retaining only the term linear in
$1/\beta J$:
\begin{eqnarray}\nonumber
\overline{G({\bf R})} = G_0({\bf R})\Bigg\{1 -2p - p \sum_{{\bf
r}\neq 0,{\bf R}} \Bigg(\frac{\pi}{8\beta J}F_1({\bf r},{\bf R})
\\\label{G2(R)_4}
+ \frac{\pi}{8\beta J(\pi-2)} F_2({\bf r},{\bf R}) \Bigg) + O((\beta J)^{-2})\Bigg\} .\quad
\end{eqnarray}
Then, using the asymptotic forms (\ref{Isc_asympt}) and (\ref{Iss_asympt}), and replacing the sum with an integral, one can show that, when $R = \sqrt{X^2 + Y^2} \to \infty$, the leading term comes from the integral which in polar coordinates reads as
\begin{eqnarray}\nonumber
&&\frac{1}{a^2} \int_{{\bf r}\neq 0, {\bf R}} d{\bf r} F_1({\bf r},{\bf R})
\\\nonumber
&&= \frac{R^2}{\pi^2} \int_{{\bf r}\neq 0,{\bf R}} \frac{rdrd\varphi}{r^2 (r^2 + R^2 - 2rR\cos\varphi)} + \ldots,
\end{eqnarray}
where the integral spans the entire system excluding areas
close to ${\bf r} = 0$ and ${\bf r} = \bf R$. This integration can be realized as follows:
\begin{eqnarray}\nonumber
\int_{{\bf r}\neq 0, {\bf R}} dr d\varphi &\to& \int_{a}^{R-a} dr \int_{0}^{2\pi} d\varphi + \int_{R+a}^{a\sqrt{N}} dr \int_{0}^{2\pi} d\varphi
\\\nonumber
&& + \int_{R-a}^{R+a} dr \int_{a/R}^{2\pi - a/R} d\varphi .
\end{eqnarray}
There is no difficulty in finding the integrals above, so, finally, one arrives at
\begin{eqnarray}\nonumber
\overline{G({\bf R})} = G_0({\bf R})\Big\{1 -2p - p
\frac{\pi}{2\pi\beta J}\ln(R/a) \Big\},
\end{eqnarray}
which can be written for small concentrations $p$ and low
temperatures $1/(\beta J)$ as
\begin{eqnarray}\label{G_s.d.(R)}
\overline{G({\bf R})} \simeq (1
-2p)\left(\frac{R}{a}\right)^{-\eta_{\mathrm{s.d.}}}
\end{eqnarray}
with
\begin{eqnarray}\nonumber
\eta_{\mathrm{s.d.}}(p) &=& \eta(0) + p\frac{T}{2J} + O(p^2) + O((T/J)^2)
\\\label{eta_s.d.}
&\simeq& \eta(0)(1+\pi p),
\end{eqnarray}
where $\eta(0)$ is the exponent of the pure system given by (\ref{eta_swa}). The factor $(1-2p)$ in (\ref{G_s.d.(R)}), which appeared naturally from the expansion, is the probability to have both sites that stand in the pair correlation function occupied with spins: $(1-p)^2 \underset{p\to 0}{\to} 1-2p$.
\section{\label{conclus}Conclusions}
The spin-spin correlation function of the $2D$ $XY$ model decays as a power law at all temperatures below the Berezinskii-Kosterlitz-Thouless transition point with a temperature dependent exponent $\eta=\eta(T/J)$. In the $2D$ $XY$ model with site or bond dilution this exponent depends on concentration $p$ of removed sites/bonds as well. The knowledge of the slope $\partial\eta/\partial p$ at point $p=0$ allows to predict the value of the exponent for small dilution concentrations: $\eta(p)\simeq\eta(0)+p(\partial\eta/\partial p)|_{p=0}$. The analytical derivation, performed here in the low-temperature limit, led to $(\partial\eta/\partial p)|_{p=0}=\pi\eta(0)$ and $2\eta(0)$ for site and bond dilution, respectively, where $\eta(0)=T/2\pi J$ is the well known result for the model without dilution. These results are illustrated in Fig. \ref{fig2}.
\begin{figure}
\center{\includegraphics[width=0.35\textwidth,angle=-90]{fig2.eps}}
\caption{\label{fig2} Analytical (lines) and Monte Carlo (squares, site dilution only) results for the ratios $\eta_{\mathrm{s.d.}}(p)/\eta(0)$ and $\eta_{\mathrm{b.d.}}(p)/\eta(0)$ ($p$ is the concentration of missing spins and bonds, respectively). Concerning the analytical results one is referred to Eqs. (\ref{eta_s.d.}) and (\ref{eta_b.d.}). The Monte Carlo data are borrowed from \cite{KapikranianEtAl07} and come from simulations with Wolff cluster algorithm at $T/J=0.08$.}
\end{figure}
The positive sign of $(\partial\eta/\partial p)|_{p=0}$ was well expected, since, as it was mentioned in Introduction, dilution can be interpreted as the increase of effective temperature. One might be tempted to equate the left sides of (\ref{eta_b.d.}) and (\ref{eta_s.d.}) to the universal value of $\eta(T_{\mathrm{BKT}})=1/4$ and identify the $T$ in the right side as the corresponding critical temperatures for site and bond dilution. Unfortunately, such an estimate of $T_{\mathrm{BKT}}(p)$ as a function of $p$ would not be quantitatively reasonable, since (\ref{eta_b.d.}), (\ref{eta_s.d.}) were obtained in the spin-wave approximation and do not hold for $T$ close to $T_{\mathrm{BKT}}(p)$.
It is worth noting that in order to compare the results for site and bond dilutions it may be more instructive to express the concentration of spinless sites, $p=$ (number of empty sites)/(number of all sites), through the actual concentration of missing bonds, $p'=$ ((four bonds)$\times$(number of empty sites))/(number of all bonds). (The latter relation holds, of course, only under the assumption of low dilution concentration, when the probability to have neighboring spinless sites is negligible.) Finally, noting that the total number of bonds in the system is two times the number of all sites, we have $p = p'/2$. Then one shall compare the exponent
\begin{equation}
\eta_{\mathrm{s.d.}}(p') = \eta(0) (1 + (\pi/2)p')
\end{equation}
and (\ref{eta_b.d.}) for $p'=p$, which means that we look at the systems with the same number of missing bonds (although in the case of site dilution all missing bonds are connected in unbreakable groups of four). One can notice that $\eta_{\mathrm{b.d.}} > \eta_{\mathrm{s.d.}}$ for the same concentration of missing bonds, which is well expected, since the disordering effect must be stronger for a completely random distribution of removed bonds in comparison to the site dilution case when removed bonds are connected in groups of four, and only these groups are distributed randomly then.
It also should be mentioned that, in principle, taking higher order terms in dilution concentration $p$ in (\ref{expansion_in_p}), one would expect to arrive at the end at the correlation function with exponent $\eta(p)$ represented by a series in powers of $p$ divergent at the percolation threshold value $p = p_{\mathrm{perc.}}$ for the square lattice [which is exactly $1/2$ for bond dilution and $\simeq 0.41$ for site dilution (see, for example, \cite{Newman_Ziff_2000})]. It is interesting in that it might give an exact value for the site percolation threshold which is not known yet. However, it might be as well not possible to carry out this calculation in an exact way, due to very high complexity.
\section{Acknowledgments}
I would like to thank Yurij Holovatch and Bertrand Berche, without whose guidance I would not start this problem in the first place, for useful discussions and corrections to the manuscript. I also acknowledge the support of the FP7 EU IRSES project N269139 'Dynamics and Cooperative Phenomena in Complex Physical and Biological Media' and the grant of the President of Ukraine for young scientists.
|
\section{Calculation of the PES and the RSES}
Our basic strategy is as follows.
Even though the set of index configurations $Z_{A},Z_{A'}$ is infinite, the rank of $\rho_{A}^{(L_{z}^{A})}$ is finite. Therefore we can find it by evaluating the rank of a suitably large submatrix which we obtain by choosing a particular set of values of $Z_A$ and $Z_{A'}$. We will use a square $d\times d$ submatrix with the same set of $Z_A$ and $Z_{A'}$, which we call $Z_{A}^{i}$ with $i\in{1,\ldots,d}$. The rank of this matrix is the same as the number of linearly independent wave functions in the set
\begin{eqnarray}
{\bf \tilde{P}} = \{ \tilde{\Phi}_i(Z_B) = \Phi(L^A_z,Z_{A}^{i},Z_B) \quad i=1,..,d \}.
\label{set_qh_1}
\end{eqnarray}
This number is equal to the rank of the matrix
\begin{eqnarray}
M_{i j} = \tilde{\Phi}_i ( Z_B^j) \qquad i=1,\ldots,d,~j=1,\ldots,d'\ge d
\label{eq:li_states}
\end{eqnarray}
where $Z_B^j = (z_{k+1},..z_N)_j \ (j=1,..,d'$) is a set of $d'\ge d$ different $(N-k)$-tuples of coordinates in the B subsystem. The rank of this matrix can be obtained (e.g.~by singular value decomposition) and equals the number of nonzero eigenvalues of $M^{\dagger}M$. In numerical calculations, all eigenvalues of $M^{\dagger}M$ will be nonzero, but a clear jump between large and small (nearly zero) eigenvalues is observed
and hence the true rank can be read off.
The scheme just described will not work without a judicious choice of
the index configurations $Z_A^i$ and $Z_B^j$. Choosing these at random will lead to very
small and greatly varying values of the matrix elements $M_{ij}$. This
induces numerical error which makes it difficult to
identify a clear cut in the spectrum of eigenvalues of $M^{\dagger}M$.
A better set of index configurations can be obtained by Monte Carlo
sampling $\psi$, but harvesting only configurations whose $Z_{A}$
satisfies a suitable constraint, which helps to select index
configurations which have good overlaps with the desired angular
momentum sector. The constraint is introduced by means of the relationship between the coordinates and the angular momentum in the Landau levels. For instance on the disk the single particle state with angular momentum $l$ has most of the probability density concentrated near a ring of radius $\sqrt{l}$. Therefore, we can assign to each Metropolis configuration $ \tilde{Z}_A = ( z_1,..,z_k )$ a $k$-tuple of angular momenta $l_A=(l_{A}^{1},..,l_{A}^k)$ by taking $l_{A}^i$ equal to the integer nearest to $|z_i|^2$. We then harvest only those $Z_A$ which satisfy $\sum^k_{i=1} l_{A}^i = L^A_z$ as index configurations.
When two configurations $Z_A$ and $Z'_A$ obtained in this way share the same $l_A$, we can discard one of them without loss of rank, because the corresponding rows of the matrix $M$ are almost exactly proportional.
Moreover, to obtain the full rank of $\rho_A^{(L_z^A)}$ it is not necessary to generate index configurations corresponding to each admissible $k$-tuple $l_A$. This is because the set of states labeled by the $Z_{A}$ is overcomplete and states with different values of $l_A$ have nonzero overlaps.
To obtain the full rank $r$ of one of the $\rho_A^{(L_z^A)}$, we typically need to take a number of indices $d$ which is only a few times larger than $r$. If we want to focus on the low lying part of the RSES, a further simplification takes place. Now the electron coordinates in the $Z_{A}^{i}$ must also lie inside the spatial region associated with subsystem $A$. E.g.~the $k$ particles in subsystem $A$ may be located in a disk of radius $r_A$, so that
$|z_i| \le r_A$ for $i \in\{1,..,k\}$. In this case the associated angular momenta $l_{A}^{i}$ must also have their probability density concentrated on rings located within this disk and the number of $l_{A}$ which satisfy the constraint that $\sum^k_{i=1} l_{A}^i = L^A_z$ is much smaller than if $z_1,\ldots,z_k$ could be located anywhere. This selection of indices allows for a very efficient evaluation of the RSES ranks in the angular momentum sectors associated with the edge, but it does make it more difficult to find the multiplicities for other values of $L_z^A$ associated with bulk excitations. To obtain those it is better to select indices $Z_{A}^{i}$ with electron coordinates which vary through all space.
Using this method, we can manage systems with up to $N=100$ and $N_A=50$ particles on a standard laptop. Note that for systems where the correspondences between the ES and the bulk and boundary excitations hold, this gives us information on the excitations for systems of up to $50$ particles, which is considerably larger than the system sizes accessible by exact diagonalization (typically no more than $20$ particles).
With more effort and resources, much larger systems should be accessible.
It should also be clear the method can be used on the plane, the cylinder and the torus.
We now present a summary of results of sample PES and RSES rank calculations. These were done for integer filling fractions on the cylinder and for the Moore-Read Pfaffian~\cite{Moore91}, Laughlin and Jain states both on the cylinder and in the spherical geometry~\cite{Haldane83}. We consider the case $N_A=N/2$ and for the RSES we define the subsystem $A$ as half of the sphere or cylinder.
For the Laughlin states at $\nu=1/2$ and $\nu=1/3$ we calculated the RSES ranks for systems of $70$ particles. The multiplicities of the low-energy edge excitations of the Laughlin state, in the thermodynamic limit, are predicted by the edge conformal field theory (CFT), which is a chiral Luttinger liquid. They are
\begin{eqnarray}
\begin{array}{|r|cccccccccc|}
\hline
\Delta L_z & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & n \\
\mbox{edge counting} & 1 & 1 & 2 & 3 & 5 & 7 & 11 & 15 & 22 & p(n)
\\ \hline
\end{array}
\label{count_Laugh}
\end{eqnarray}
Here $\Delta L_z$ is the relative angular momentum between the quasihole excitations and the Laughlin groundstate and $p(n)$ is the number of partitions of
$n$.
We have computed the RSES ranks up to $L^A_z = 20$ and find that for both $\nu=1/2$ and $\nu=1/3$, they match the edge counting above.
\begin{figure}[thb]
\begin{center}
\includegraphics[width= 7.5cm, height=5cm]{M_energies_plot_latest.pdf}
\caption{Plot of the spectra of the matrices $M^{\dagger}M$ used in the calculation of the RSES ranks for a $\nu=\frac{1}{2}$ Laughlin state at $N=70$ and $N_A=35$ on a sphere. The negative logarithms of the eigenvalues for each $L_{z}^{A}$ block of $\rho_A$ are plotted (the largest eigenvalue is normalized to $1$ at each $L_{z}^{A}$). A clear gap is visible for each value of $L_{z}^{A}$ and the counting of large eigenvalues (at low $\epsilon$) matches the partition numbers.}
\label{fig:Laughlin13}
\end{center}
\end{figure}
Fig.~\ref{fig:Laughlin13} shows the spectra of the matrices $M^{\dagger}M$
for the Laughlin state at $\nu=1/2$ at different values of $L_{z}^{A}$, in the way in which entanglement spectra are usually presented ($L^A_z$ is given relative to the state with the lowest $L^A_z$, analogously to $\Delta L_z$ above). A clear gap is visible for each value of $L_{z}^{A}$ and the counting of large eigenvalues matches the partition numbers. Of course this plot is not a plot of the true RSES. Repeats of the RSES rank calculation with different initialization of the random number generator will yield different, but similar looking plots.
For the bosonic and fermionic MR states, we also computed the RSES at $N=70$ and $L^A_z\le 20$, obtaining the expected CFT countings~\cite{Milovanovic96}. We also checked that, by changing $N_A$ or by including localized quasiholes, one may obtain the CFT countings for different topological sectors, in analogy to the results for the OES\cite{Li08,Papic11}.
We have computed the PES ranks beyond the universal CFT numbers, for $\nu=1$, $\nu=\frac{1}{2}$, and $\nu=\frac{1}{3}$ and for the fermionic and bosonic MR states, for many values of $L^A_z$ at $N=50$ and for all $L^A_z$ in smaller systems,
obtaining the expected finite size countings~\cite{Read96,Hermanns11}.
At integer filling $\nu\in\mathbb{N}$ the OES ranks are trivial but the RSES ranks are the full ranks of the angular momentum sectors of the Hilbert spaces for $N_{A}$ particles in $\nu$ Landau levels.
For $\nu=1$, the ranks are just the partition numbers (\ref{count_Laugh}) in the thermodynamic limit, in agreement with Ref.~\cite{Rodriguez09}, which treated the $\nu=1$ system analytically in a second quantized formulation. For general integer filling $\nu=p$, the ranks depend on the residue of $N_A$ modulo $p$. E.g.~for $\nu=2$, a generating function for the thermodynamic countings is
\begin{equation}
\label{eq:nu2count}
Z_{\nu=2} = \left(\sum_{m\in\mathbb{Z}} q^{m^2-s m}\right)\left(\prod_{k > 0} \frac{1}{1 - q^k}\right)^2,
\end{equation}
where $s=N_{A} (\mathrm{mod}~2)$.
Similar formulas may be obtained for higher integer $\nu$. In fact, it can be shown that
for $N_A=0~({\rm mod}~\nu)$ the ranks always equal the numbers of $\nu$-colored generalized Frobenius partitions~\cite{Andrews84}.
We checked that our method reproduces these numbers for $\nu\le 3$, for systems of around $30$ particles and necessarily for modest angular momenta, as the ranks grow quickly.
\begin{figure}[hbt]
\begin{center}
\vspace*{-7mm}
\includegraphics[width= 8.5cm, height=5cm]{RSES_JainN20-22.PDF}
\caption{Plot of the spectra of the matrices $M^{\dagger}M$ used in the calculation of the RSES ranks for a $\nu=\frac{2}{5}$ Jain state on a sphere at $N=20$ and $N_A=10$ (left) and at $N=22$ and $N_A=11$ (right). Negative values of $\epsilon$ occur because we used an unnormalized Jain state (normalization shifts the vertical axis by a constant). The counting $1,2,5,10,\ldots$ of the branches is consistent with a pair of noninteracting Luttinger liquids. Differences in total counting for even vs.~odd $N_A$ arise analogously to those for $\nu=2$ (cf. Eq.~(\ref{eq:nu2count})).}
\label{fig:Jain25}
\end{center}
\end{figure}
Finally, we apply our method to the Jain state at $\nu=\frac{2}{5}$.
Jain states~\cite{Jain89} have so far produced serious challenges in ES calculations because they lacked a clear `entanglement gap' at the accessible sizes. While the numerical method described here calculates RSES ranks, we can expect the low lying parts of the spectra of the matrices $M^{\dagger}M$ to converge to yield the low lying parts of the true RSES, as the size of the matrices $M$ is increased. Fig.~\ref{fig:Jain25} shows the spectra of the matrices $M^{\dagger}M$ at low $L_{z}^{A}$ for the $\nu=\frac{2}{5}$ CF state at $N=20$ and $N=22$, obtained with matrices $M$ which are large enough to allow the pattern of entanglement energies to emerge. Note that so far, no ES for this state has been published for $N>10$, due to the difficulty of obtaining the wave function in angular momentum space at large $N$. At low $L_z^A$ values, we observe several branches of edge states in the RSES. Each branch has ranks $1,2,5,10,\ldots$, consistent with the counting for a pair of non-interacting Luttinger liquid edges. The total rank at each $L_z^A$ equals the dimension of the spaces of CF states with $N_A$ CFs in $2$ CF Landau levels at the given value of $L_{z}^{A}$. The ranks are in general smaller than those for $\nu=2$ because some CF states disappear in the lowest LL projection. Note that the CF LLs we used for the $A$ system are the same as those for the full system. This is natural for the low $L_{z}^{A}$ edge states of the RSES. For a description of the full PES, one should consider CFs
at a higher effective flux,
see Ref.~\onlinecite{Sterdyniak11_PRL} for details.
We further checked that the space of eigenvectors of $M^{\dagger}M$ at each $L_z^A$ has very close to unit overlap with the space spanned by the vectors $v_i=\phi^{CF}_{i}(Z^{j}_{A})$, where the $\phi^{CF}_{i}$ are the CF trial wave functions for system $A$ and the $Z^{j}_{A}$ are the configurations of the particles in subsystem $A$ that were used in the construction of $M$.
It appears that the low lying RSES of this CF state is excellently described by the CF trial wave functions for the edge excitations of the $A$ system.
In future work, we intend to expand our results on CF states and on the calculation of entanglement energies.
\noindent \textbf{Acknowledgments:} The authors thank Nicolas Regnault for illuminating discussions. SHS and JKS acknowledge the support and hospitality of the Aspen Center for Physics. JKS and IDR were supported by SFI Principal Investigator award 08/IN.1/I1961. SHS is supported by EPSRC grant EP/I032487/1.
\noindent \textbf{Note added:} As this work was being completed, we became aware of research studying the RSES by A.~Sterdyniak, A.~Chandran, N.~Regnault, B.~A.~Bernevig and P.~Bonderson~\cite{Sterdyniak11_arxiv}, and by J.~Dubail, N.~Read and E.~H.~Rezayi~\cite{Dubail11}.
|
\section{Introduction}
A differentiable manifold $(M^{2n}, {\omega}, \theta)$ provided with a non-degenerate 2-form ${\omega}$ and a closed 1-form $\theta$ is called a locally conformal symplectic (l.c.s.) manifold, if $d{\omega} = -{\omega} \wedge \theta$, $d\theta = 0$. The 1-form $\theta$ is called the Lee form of ${\omega}$.
The class of l.c.s. manifolds has attracted
strong interests among geometers in recent years. For instance, Vaisman showed that l.c.s. manifolds
may be viewed as phase spaces for a natural generalization of Hamiltonian dynamics \cite{Vaisman1985}. Bande and Kotschick showed that a pair composed of a contact manifold and a contactomorphism is naturally associated with a l.c.s. manifold \cite{BK2010} (see also Proposition \ref{mtc} and Proposition \ref{inv2} below). Furthermore, l.c.s. manifolds together with contact manifolds are the only transitive Jacobi manifolds \cite[Remark 2.10]{Marle1989}. It is also worth mentioning that locally conformal K\"ahler manifolds, a natural subclass of l.c.s. manifolds, are actively studied in complex geometry, e.g. see \cite{HK2011}, \cite{OV2010}.
Note that a l.c.s. manifold is locally conformal equivalent to a symplectic manifold, i.e. locally $\theta = df$ and $ {\omega} = e^{-f} {\omega}_0$, $d{\omega}_0 = 0$. By the Darboux theorem all symplectic manifolds of the same dimension are locally equivalent. Hence symplectic manifolds have only global invariants, and cohomological invariants are most natural among them. First (co)homological symplectic invariants were proposed in works by Gromov and Floer then followed by works by McDuff, Hofer and Salamon, Fukaya and Ono, Ruan, Tian, Witten and many others, including the first author of this note. This approach was based on the use of the theory of elliptic differential operators with purpose to make regular certain Morse (co)homology theory or the intersection theory on the infinite dimensional loop space on a symplectic manifold $M^{2n}$, or on the space of holomorphic curves on $M^{2n}$. This elliptic (co)homology theory has huge success, but the computational part of the theory is quite complicated. Almost at the same time, a ``linear" symplectic cohomology theory has been developed, beginning with the paper by Brylinski \cite{Brylinski1988}, followed by Bouche \cite{Bouche1990}, and then by other peoples (see \cite{FIL1998}, \cite{PV2006}, \cite{TY2009}). This theory is mostly motivated by
analogues in K\"ahler geometry, the Dolbeault theory, and the cohomology theory for differential equations developed by Vinogradov and his school.
This linear symplectic cohomology theory has not yet drawn as much attention as it, to our opinion, should have. This is, probably, due to the fact that its potentially important applications are still in a phase of elaboration. The computational part of the linear theory seems to be not so complicated as in the elliptic theory, and this is an advantage of its.
In our paper we further develop the linear symplectic cohomology theory and extend it to l.c.s. manifolds. This is possible due to the validity of the Lefschetz
decomposition for these manifolds. The main tool is the spectral sequence
developed by Di Pietro and Vinogradov for symplectic manifolds, which has
been now adapted and developed further for l.c.s. manifolds. We obtain new results and applications
even for symplectic manifolds. In
particular we unify various isolated results of the linear symplectic
cohomology theory.
The structure of this note is as follows. In section 2 we introduce important linear operators on
l.c.s. manifolds and study their properties. The Lefschetz filtration on the space ${\Omega} ^* (M^{2n})$ of a l.c.s. manifold $(M^{2n}, {\omega}, \theta)$ is discussed in section 3 together with differential operators respecting this filtration. Then we use this filtration to construct primitive
cohomology groups for $(M^{2n}, {\omega}, \theta)$ (Definition \ref{newcomg}). Some simple properties of these groups are fixed in Proposition \ref{plus1}, Proposition \ref{primi} and Proposition \ref{dpm1}, and their relations with previously proposed constructions are discussed (Remark \ref{his1}).
The spectral sequence associated with the Lefschetz filtration is studied in section 4. In particular, its $E_1$-term is compared with the primitive (co)homological groups (Lemma \ref{fk1}) and the conformal invariance of this term is proven (Theorem \ref{conf}). In section 5 we find some cohomological conditions on $(M^{2n}, {\omega}, \theta)$
under which this spectral sequence stabilizes at the $E_{t}$-term (Theorems \ref{stab1}, \ref{nil}, \ref{stcom}). The last of these theorems gives an answer to the Tseng-Yau question on relations between
the primitive cohomology and the de Rham cohomology of a compact symplectic manifold.
In section 6 we specialize the previous theory to K\"ahler manifolds and prove that for K\"ahler
manifolds the spectral sequence stabilizes already at its first term (Theorem \ref{kaehler1}). In section 7 we compute the primitive cohomology groups of a compact $(2n+2)$-dimensional l.c.s. nilmanifold and a compact 4-dimensional l.c.s. solvmanifold (Propositions \ref{mil1}, \ref{solv1}). We study some properties of primitive cohomology groups of l.c.s. manifolds associated with a co-orientation preserving contactomorphism (Proposition \ref{inv2}). In particular, we show that the compact l.c.s. solvmanifold is associated with a non-trivial co-orientation preserving contactomorphism (Theorem \ref{new1}).
The cohomological theory developed in this note and its analogues have a much wider area of applications. For instance, it may be naturally adopted to the class of Poisson symplectic stratified spaces introduced in \cite{LSV2010}, since these singular symplectic spaces also enjoy the Lefschetz decomposition.
This project was started as a joint work of us with Alexandre Vinogradov based on H.V.L. preliminary results on l.c.s. manifolds. Alexandre Vinogradov has suggested
us to extend the results to a slightly larger category of twisted symplectic manifolds. He made considerable contributions to improve the original text written by H.V.L., which we appreciate very much. Eventually we have noticed that our viewpoints are so different, so we decide to write the subject separately: in this paper we deal only with l.c.s. manifolds and Alexandre Vinogradov will deal with the extension to twisted symplectic manifolds.
\section{Basic operators on a l.c.s. manifold}
In this section we introduce and study basic linear differential operators acting on differential forms on a
l.c.s. manifold $(M^{2n}, {\omega}, \theta)$.
The first operator we need is the Lichnerowicz deformed differential $d_\theta: {\Omega}^* (M^{2n}) \to {\Omega}^*(M^{2n})$,
\begin{equation}
d_\theta (\alpha): =d\alpha + \theta \wedge\alpha.
\label{lic1}
\end{equation}
Clearly $d_\theta ^2 = 0$ and $ d_\theta ({\omega}) = 0$. The resulting Lichnerowicz cohomology groups (also called the Novikov cohomology groups) are important conformal invariants of l.c.s. manifolds.
Recall that two l.c.s. forms ${\omega}$ and ${\omega}'$ on $M^{2n}$ are {\it conformally equivalent}, if
${\omega}' = \pm (e^ f) {\omega}$ for some $f \in C^\infty (M^{2n})$. In this case the corresponding Lee forms $\theta$ and $\theta '$ are cohomologous: $\theta' = \theta \mp df$, hence
$ d_{\theta}$ and $d_{\theta'}$ are {\it gauge equivalent}:
$$d_{\theta'} ( \alpha) = (d_{\theta} \mp df \wedge )\alpha = e ^{\pm f} d (e ^{\mp f} \alpha).$$
It follows that $H^* ({\Omega}^* (M^{2n}), d_\theta)$ and
$H^*({\Omega}^*(M^{2n}), d_{\theta'})$ are isomorphic. The isomorphism $ I_f : H^* ({\Omega}^* (M^{2n}), d_{\theta}) \to H^*({\Omega}^*(M^{2n}), d_{\theta'})$ is given by the conformal transformation $[\alpha] \mapsto [\pm e^ f \alpha]$.
Note that $d_\theta$ does not satisfy the Leibniz property, unless $\theta = 0$, since
\begin{equation}
d_{\theta} ( \alpha \wedge \beta) = d_{\theta} \alpha \wedge \beta + (-1) ^{deg\, \alpha}\alpha \wedge d\beta = d\alpha \wedge \beta +(-1) ^{deg\, \alpha}\alpha \wedge d_\theta \beta.
\label{lic2n}
\end{equation}
Thus the cohomology group $H^*({\Omega}^* (M^{2n}), d_{\theta})$ does not have a ring structure, unless $\theta = 0$.
The formula (\ref{lic2n}) also implies that $H^*({\Omega}^* (M^{2n}), d_{\theta})$ is a $H^*(M, {\mathbb R})$-module.
Now let us consider the next basic linear operator
\begin{equation}
L: {\Omega}^* (M^{2n}) \to {\Omega}^* (M^{2n}) , \: \alpha \mapsto {\omega} \wedge \alpha. \label{defl}
\end{equation}
Substituting $\alpha := {\omega}$ in (\ref{lic2n}) we obtain a nice relation between $d$, $L$ and $d_{\theta}$
\begin{equation}
d_\theta L = Ld.\label{lic3}
\end{equation}
The identity (\ref{lic3}) suggests us to consider a family of operators $d_{k\theta}$, which we abbreviate as
$d_k$ if no misunderstanding occurs. We derive immediately from (\ref{lic3})
\begin{equation}
d _k L^ p = L^p d _{k -p}. \label{lic4}
\end{equation}
The following Lemma is a generalization of (\ref{lic2n}) and it plays an important role in our study of the spectral sequences introduced in later sections. It is obtained by straightforward calculations, so we omit its proof.
\begin{lemma}\label{comd1} For any $\alpha, \beta \in {\Omega}^* (M^{2n})$ we have
\begin{equation}
d_{k+l} (\alpha \wedge \beta) = d_k \alpha \wedge \beta + (-1) ^{ deg\, \alpha}\alpha \wedge d_l \beta.\label{comd2}
\end{equation}
Consequently
\begin{equation}
d_k \alpha \wedge d_l \beta = d_{k+l} ( \alpha \wedge d_l \beta).\label{comd3}
\end{equation}
Formula (\ref{comd2}) yields the induced map $H^*({\Omega}^*(M^{2n}), d_k) \times H^* ({\Omega}^* (M^{2n}), d_l) \to H^* ({\Omega}^*(M^{2n}), d_{k+l})$.
\end{lemma}
Denote by $G_{{\omega}}$ the section of the bundle $\Lambda ^2 TM^{2n}$ such that for all $x \in M^{2n}$ the linear map $i_{G_{\omega} (x)} : T^*_x M^{2n} \to T_xM^{2n}, \, V \mapsto i_V (G_{\omega} (x)),$ is the inverse of the map $I_{\omega} : T_x M^{2n} \to T^*_xM^{2n},\, V \mapsto i_V {\omega}$.
Clearly $G_{\omega}$ defines a bilinear pairing: $T^*M^{2n} \times T^*M^{2n}\to C^\infty (M^{2n})$. Denote by $\Lambda ^p G_{\omega}$ the associated pairing: $\Lambda^p (T^*M) \times \Lambda ^p(T^*M) \to C^\infty (M^{2n})$. The l.c.s. form ${\omega}$ and the associated bi-vector field $G_{\omega}$ define {\it a l.c.s. star operator} $*_{\omega}: {\Omega}^p (M^{2n}) \to {\Omega} ^{2n-p}(M^{2n})$ as follows \cite[\S 2.1]{Brylinski1988}.
\begin{equation}
*_{\omega} : {\Omega} ^p (M^{2n}) \to {\Omega}^{2n-p} (M^{2n}),\, \beta \wedge *_{\omega} \alpha : = \Lambda ^p G_{\omega} (\beta, \alpha) \wedge \frac {{\omega} ^n} { n !},
\label{sst}
\end{equation}
for all $\alpha, \beta \in {\Omega} ^p (M^{2n})$.
Using \cite[Lemma 2.1.2]{Brylinski1988} we get easily
\begin{equation}
*_{\omega} ^2 = Id.\label{bryss1}
\end{equation}
We define {\it the l.c.s. adjoint} $L^*$ of $L$ and { \it the l.c.s. adjoint} $(d_k)^*_{\omega}$ of $d_k$ with respect to the l.c.s. form ${\omega}$ as follows
\begin{equation}
L^* : {\Omega} ^{p} (M^{2n})\to {\Omega}^{p-2} (M^{2n}), \, \alpha^p \mapsto - *_{\omega} L *_{\omega} \alpha^p. \label{defls}
\end{equation}
\begin{equation}
(d_k)^*_{\omega}: {\Omega} ^{p} (M^{2n})\to {\Omega}^{p-1} (M^{2n}), \, \alpha^p \mapsto (-1) ^p * _{\omega} d_{n+k-p} *_{\omega} (\alpha^p). \label{defnbls}
\end{equation}
For symplectic manifolds our definition of $(d_k)^*_{\omega}$ agrees with the one in \cite[\S 1]{Yan1996}, it is different from the one in \cite[Theorem 2.2.1]{Brylinski1988} by
sign (-1).
A section $g $ of the bundle $S^2 T^*M^{2n}$ is called {\it a compatible metric}, if there is an almost complex structure $J$ on $M^{2n}$ such that $g (X, Y) = {\omega}(X, JY)$. In this case $J$ is called {\it a compatible almost complex structure.}
Recall that the Hodge operator $*_g$ is defined as follows
\begin{equation}
*_g : {\Omega} ^p (M^{2n}) \to {\Omega}^{2n-p} (M^{2n}),\, \beta \wedge *_g \alpha : = \Lambda ^p G_g (\beta, \alpha) \wedge \frac {{\omega} ^n} { n !},\label{hodgets}
\end{equation}
where $G_g \in \Gamma (S^2 TM^{2n})$ is the ``inverse of $g$", i.e. it is defined in the same way as we define $G_{\omega}$ above: for all $x \in M^{2n}$ the linear map $i_{G_g (x)} : T^*_x M^{2n} \to T_xM, \, V \mapsto i_V (G_g (x)),$ is the inverse of the map $I_g : T_x M \to T^*_xM,\, V \mapsto i_V (g)$. We also denote by $\Lambda ^p G_g$ the associated pairing: $\Lambda^p (T^*M) \times \Lambda ^p(T^*M) \to C^\infty (M^{2n})$ induced by $G_g$, (see also \cite[p.105]{Brylinski1988} for comparing $\Lambda ^p G_{\omega}$ with $\Lambda^p G_g$).
Using \cite[Lemma 5.5]{Voisin2007} we get easily
\begin{equation}
*_g ^ 2 (\alpha ^p) = (-1) ^p\alpha ^p \text{ for } \alpha ^p \in {\Omega} ^p (M^{2n}). \label{voisin5.5}
\end{equation}
\begin{lemma}\label{fund} 1. The space of metrics compatible with a given l.c.s. form ${\omega} \in {\Omega}^2 (M^{2n})$ is contractible.
2. (cf. \cite[chapter II, 6.2.1]{Voisin2007}) In the presence of a compatible metric $g$ on $M^{2n}$ we have
\begin{equation}
L^* = \Lambda \label{lel}
\end{equation}
where $\Lambda = (*_g)^{-1} L *_g$ is the adjoint of $L$ with respect to the metric $g$.
\end{lemma}
\begin{proof} 1. The proof for the first assertion goes in the same way as for the case of symplectic manifolds, so we omit its proof.
2. The second assertion of Lemma \ref{fund} is a simple consequence of the following
\begin{lemma}\label{hk}\cite[Theorem 2.4]{Brylinski1988} Assume that $(M^{2n}, J, g)$ is an almost Hermitian manifold and ${\omega}$ is the associated almost symplectic form. For $\alpha \in {\Omega} ^{p, q}(M^{2n})$ we have
$$*_{\omega} (\alpha) = \sqrt{-1}^{p-q} *_g (\alpha).$$
Here we extend $*_{\omega}$ and $*_g$ ${\mathbb C}$-linearly on ${\Omega} ^* (M^{2n})\otimes {\mathbb C}$.
\end{lemma}
This completes the proof of Lemma \ref{fund}.
\end{proof}
Let $\pi_{k} : {\Omega}^* (M^{2n}) \to {\Omega}^k (M^{2n})$ be the projection. Denote $\sum _{i=0} ^{2n} (n-k)\pi_{k}$ by $A$.
Using well-known identities in K\"ahler geometry for $(\Lambda, L, A)$, see e.g. \cite[(IV), chapter I]{Weil1958}, \cite[p.121]{GH1978}, \cite[Lemma 6.19]{Voisin2007}, Lemma \ref{fund} implies immediately the following
\begin{corollary}\label{hks} (cf. \cite[\S 1]{Lychagin1979}, \cite[Corollary 1.6]{Yan1996}) On any l.c.s. manifold $(M^{2n}, {\omega}, \theta)$ we have
\begin{equation}
L ^* = i (G_{\omega}),\label{lss}
\end{equation}
\begin{equation}
[L ^*, L] = A,\: [A, L] = - 2L, \: [A, L^*]= 2L^*.\label{sl2}
\end{equation}
\end{corollary}
The relation in (\ref{sl2}) shows that $(L^*, L, A)$ forms a ${\mathfrak sl}_2$-triple, which has many important consequences
for l.c.s. manifolds.
\begin{proposition}\label{comdn} The following commutation relation hold
\begin{equation}
L^* (d _k)^*_{{\omega}}= (d _{k-1})^*_{\omega} L^*.\label{lic4s}
\end{equation}
\end{proposition}
\begin{proof} Clearly (\ref{lic4s}) is obtained from (\ref{lic4}) by applying the LHS and RHS of (\ref{lic4}) the l.c.s. star operator on the left and on the right, taking into account (\ref{bryss1}).
\end{proof}
\section{Primitive forms and primitive cohomologies}
In this section we introduce the notions of primitive forms and coeffective forms on a l.c.s. manifold
$(M^{2n}, {\omega}, \theta)$, using the linear operators $L$ and $L^*$ defined in the previous section. As in the symplectic case we obtain a Lefschetz decomposition of the space ${\Omega}^*(M^{2n})$ induced by
primitive forms and coeffective forms together with various linear differential operators respecting
this decomposition as well as an associated filtration (Propositions \ref{clo2} and \ref{hlc}).
The natural splitting of the introduced differential operators according to the Lefschetz decomposition leads to new cohomology groups
of $(M^{2n},{\omega}, \theta)$ (Definition \ref{newcomg}). In Propositions \ref{plus1}, \ref{primi}, \ref{dpm1} we fix simple properties of these new cohomology groups.
At the end of this section we compare our construction with related constructions in \cite{Lychagin1979}, \cite{Bouche1990}, \cite{Rumin1994}, \cite{FIL1996}, \cite{Yan1996}, \cite{PV2006}, \cite{Pietro2006}.
\begin{definition}\label{primco} (\cite{Bouche1990}, \cite{Yan1996}, cf. \cite{Weil1958}, \cite{GH1978}) An element $\alpha \in \Lambda ^k T^*_xM^{2n}$, $0\le k \le n$, is called {\it primitive} (or {\it effective}), if
$L^{n-k+1} \alpha = 0$. An element $\alpha \in \Lambda ^k T^*_xM^{2n}$, $n+1\le k \le 2n$, is called {\it primitive}, if $\alpha = 0$. An element $\beta \in \Lambda^k T^*_x M^{2n}$ is called {\it coeffective}, if
$L \beta = 0$.
\end{definition}
\begin{remark}\label{inv1} 1. Wells in \cite{Wells1986} refers to Lefschetz \cite{Lefschetz1924} and Weil \cite{Weil1958} for the terminology ``Lefschetz decomposition" and ``primitive forms". Many
mathematicians prefer ``Lepage decomposition" and ``effective forms" following Lepage in \cite{Lepage1946}.
2. Clearly the notion of primitive form as well as the notion of coeffective form
depends only on the conformal class of a l.c.s. form ${\omega}$.
\end{remark}
The relation (\ref{sl2}) between linear operators $L, L^*$ and $A$ leads to Lemma \ref{yanpr} below characterizing primitive forms and coeffective forms. The resulting Lefschetz decomposition of the space $\Lambda T^* M^{2n}$
is a direct consequence of the ${\mathfrak sl}(2)$-module theory. Various variants of Lemma \ref{yanpr} for symplectic manifolds appeared in many works, beginning possibly with a paper by Lepage \cite{Lepage1946}, with later applications in K\"ahler geometry \cite{Weil1958}, \cite{GH1978}, \cite{Voisin2007}, in a theory of second-order differential equations \cite{Lychagin1979}, in symplectic geometry \cite{Bouche1990}, \cite{Yan1996}, etc..
We denote by $P^k_x(M^{2n})$ the set of primitive elements in $\Lambda^k T^*_xM^{2n}$.
\begin{lemma}\label{yanpr} 1. An element $\alpha \in \Lambda ^k T^*_xM^{2n}$ is primitive, if and only if
$L^* \alpha = 0$.\\
2.
An element $\beta \in \Lambda ^k T^*_xM^{2n}$ is coeffective, if and only
if $*_{\omega} \beta $ is primitive.\\
3. We have the following Lefschetz decomposition for $n \ge k \ge 0$:
\begin{equation}
\Lambda ^{n-k}T^*_x M^{2n}= P^{n-k }_x (M^{2n}) \oplus LP^{n-k-2}_x(M^{2n}) \oplus L^2 P^{n-k-4}_x(M^{2n}) \oplus \cdots,
\label{lh1}
\end{equation}
\begin{equation}
\Lambda ^{n+k}T^*_x M^{2n}=L^k P^{n-k }_x (M^{2n}) \oplus L^{k+1}P^{n-k-2}_x(M^{2n}) \oplus \cdots .
\label{lh2}
\end{equation}
\end{lemma}
From Lemma \ref{yanpr} we get immediately
\begin{corollary}\label{inj1}
1. $L^k : \Lambda^{n-k} T^*_xM^{2n} \to \Lambda ^{n+k} T^*_xM^{2n}$ is an isomorphism, for $0\le k \le n$.\\
2. $L : \Lambda^{n-k-2} T^*_xM^{2n}\to \Lambda^{n-k}T^*_xM^{2n}$ is injective, for $k =-1, 0, 1, \cdots , n-2$.
\end{corollary}
It is useful to introduce the following notations.
Denote by $P^{n-k}M^{2n}$ the subbundle in $\Lambda T^*M^{2n}$ whose fiber is $P^{n-k} _x (M^{2n})$.
Let ${\mathcal P}^{n-k}(M^{2n})\subset {\Omega}^{n-k} (M^{2n})$ be the space of all smooth $(n-k)$-forms with values in $P^{n-k}M^{2n}$.
Elements of ${\mathcal P}^{n-k}(M^{2n})$
are called {\it primitive $(n-k)$-forms}. Let us set (cf. \cite{TY2009})
\begin{equation}
\Ll ^{s, r} : = L^s {\mathcal P}^r \text { for } 0 \le s, r \le n. \label{lty1}
\end{equation}
Put ${\mathcal P}^* (M^{2n}): = \oplus _r {\mathcal P} ^r (M^{2n})$. Then Lemma \ref{yanpr} yields the following decompositions, which we call {\it the first and second Lefschetz decompositions}
\begin{equation}
{\Omega}^*(M^{2n}) ={\mathcal P}^*(M^{2n}) \oplus L {\Omega}^* (M^{2n}) = \bigoplus _{0\le 2s + r \le 2n} \Ll ^{s, r}.\label{lty2}
\end{equation}
Now we consider the interplay between the Lefschetz decompositions
(\ref{lty2}) and the linear differential operators introduced in the previous section.
Iterating the action of $L$ on $K^*: = {\Omega} ^* (M^{2n})$, we define the following filtration
\begin{equation}
F^0 K^*: = K ^* \supset F^1 K^*: = LK^* \supset \cdots
\supset F^k K^* :=L^{k} K^*\supset \cdots \supset F^{n+1} K^* =
\{ 0\} .
\label{vin2}
\end{equation}
\begin{proposition}\label{clo2} 1. The subset $F^k K^* $ is stable with respect to $d_p$ for all $k$ and $p$.\\
2. For any $\gamma \in {\Omega} ^1 (M^{2n})$ we have
$$\gamma \wedge \Ll^{0, n-k} \subset \Ll^{0, n-k+1}\oplus \Ll^{1,n-k-1}.$$
\end{proposition}
\begin{proof} 1. The first assertion of Proposition \ref{clo2} follows from the identity $d_p ({\omega} ^k \wedge \phi) = {\omega}^k\wedge d_{p-k} \phi$ for $\phi \in {\Omega}^*(M^{2n})$.\\
2. Assume that $\alpha \in {\mathcal P}^{n-k}(M^{2n}) = \Ll^{0,n-k}$. Then $L^{k +1} (\gamma \wedge \alpha) = \gamma \wedge L^{k+1} (\alpha) = 0$. Taking into account the decomposition of $\gamma\wedge \alpha$ according to the second Lefschetz decomposition we obtain the second assertion of Proposition \ref{clo2} immediately.
\end{proof}
\begin{remark}\label{rem:left1} 1. The relation $d_p({\omega} ^k \wedge \phi) = {\omega} ^ k \wedge d_{p-k} \phi$ can be also interpreted as an interplay between different filtered complexes $(F^*K^*, d_k)$ and
$(F^*K^*, d_p)$. We shall investigate this interplay deeper in the next section.
2. We observe that the decompositions (\ref{lh1}), (\ref{lh2}) and (\ref{lty2}) are compatible with the filtration (\ref{vin2}) in the following sense. For any $p\ge 0$ and $0 \le k \le 2n$ we have
\begin{equation}
F^p K^* \cap {\Omega} ^{k} (M^{2n}) = \oplus_{i =0}^{[{k\over 2}]-p} \Ll^{p+i, k -2p-2i} \text{ if } k \ge 2p,
\label{fil1}
\end{equation}
\begin{equation}
F^p K^* \cap {\Omega} ^{k} (M^{2n}) = 0
\text{ if } k < 2p,
\label{fil1a}
\end{equation}
The decomposition in (\ref{fil1}) and (\ref{fil1a}) will be called {\it the induced Lefschetz decomposition}. It is important for understanding the spectral sequences introduced in the next section.
\end{remark}
\begin{proposition}\label{hlc} The following inclusions hold
\begin{equation}
d_r\Ll^{p, q-p} \subset \Ll^{p,q-p+1}\oplus \Ll^{p+1,q-p-1},
\label{prdlt}
\end{equation}
\begin{equation}
(d_r)^*_{\omega} {\mathcal P}^{n-k}(M^{2n})\subset {\mathcal P}^{n-k-1} (M^{2n}).
\label{prde1}
\end{equation}
\end{proposition}
\begin{proof} Let $\beta\in {\mathcal P}^{q} (M^{2n})= \Ll ^{0,q}$, so $L^{n-q+1}\beta = 0$. We derive from (\ref{lic4})
\begin{equation}
L^{n-q+1} d _r\beta = d_{r+n - q +1}L^{n-q+1} \beta = 0.
\end{equation}
Using (\ref{lh1}) and (\ref{lh2})
we get $d_{r} \beta \in {\mathcal P}^{q+1}(M^{2n}) + L{\mathcal P}^{q-1}(M^{2n})$. This proves the inclusion (\ref{prdlt}) of Proposition \ref{hlc} for $p = 0$.
The inclusion (\ref{prdlt}) for $p \not = 0$ follows from the particular case $p = 0$ and the identity $d_r L^p = L^p d_{r-p}$.
Assume that $\beta \in {\mathcal P}^{n-k} (M^{2n})$. Taking into account (\ref{lic4s}) we obtain
$$ L^*(d_r)^*_{\omega} \beta =(d _{r-1})^*_{\omega} L ^* \beta,$$
which is zero since $\beta$ is primitive.
Hence $(d_r)^*_{\omega} \beta$ is also primitive. This proves (\ref{prde1}) and completes the proof of Proposition \ref{hlc}.
\end{proof}
Now we will show several consequences of Proposition \ref{hlc}. Denote by $\Pi_{pr}$ the projection ${\Omega}^* (M^{2n}) \to {\mathcal P}^*(M^{2n})$ according to the Lefschetz decomposition in (\ref{lty2}).
Set
$$d_k^+:= \Pi_{pr}d_k.$$
Using the first Lefschetz decomposition and Proposition \ref{hlc} we decompose the operator $d_k: {\Omega} ^q (M^{2n}) \to {\Omega} ^{q +1} (M^{2n})$ for $ 0\le q \le n$ as follows (cf. \cite{TY2009}).
\begin{equation}
d_k = d_{k}^+ + L d_{k}^-,\label{hoya}
\end{equation}
where $d^-_{k} : {\Omega} ^q (M^{2n}) \to {\Omega} ^{q-1} (M^{2n}), \, 0\le q \le n$. Note that $ d ^-_k$ is well-defined, since $L : {\Omega} ^{q-1} (M^{2n}) \to {\Omega} ^{ q+1}(M^{2n})$ is
injective. It is straightforward to check
\begin{equation}
d_k^+(\Ll ^{s, r}) = 0 \text { if } s\ge 1, \text{ and } d_{k}^-(\Ll ^{s, r}) \subset \Ll ^{s, r-1}.\label{lty3}
\end{equation}
\begin{lemma}\label{hoya1} (cf. \cite[Lemma 2.5, II]{TY2009}) The operators $d_k^+$, $d_{k-1}^-$ satisfy the following properties
\begin{eqnarray} (d_k^+) ^2 (\alpha^q) = 0,\label{com0}\\
d_{k-1} ^-d_k^-(\alpha ^q)= 0 , \text { if } q \le n, \label{com1}\\
(d_k^- d_k^+ + d_{k-1}^+ d^-_k) \alpha ^ q = 0, \text{ if } q \le n-1, \label{com1a}\\
(d_{k-1}) ^* _{\omega} (d_k) ^* _{\omega} (\alpha^q) = 0.\label{prag0510}
\end{eqnarray}
\end{lemma}
\begin{proof} We use the equality $d_k^2=0$ in the form $d_k(d_k^++Ld_k^-)=0$. Using (\ref{lic4}) we get
$$
(d_k^+)^2+L(d_k^-d_k^++d_{k-1}^+d_k^-)+L^2d_{k-1}^-d_k^-=0.
$$
Now taking into account (\ref{lty3}) and the injectivity of the operators $L : {\Omega} ^{q} (M^{2n}) \to {\Omega} ^{q+2}(M^{2n})$ and $L ^2 : {\Omega} ^{q-1} (M^{2n}) \to {\Omega} ^{q+3} (M^{2n})$ for $q \le n-1$ we obtain (\ref{com0}), (\ref{com1}), and (\ref{com1a}).
Finally, (\ref{prag0510}) is a consequence of $d_k^2=0$ and $*_\omega^2=Id$.
\end{proof}
\medskip
Proposition \ref{hlc} and Lemma \ref{hoya1} lead to new cohomology groups associated with a l.c.s. manifold $(M^{2n}, {\omega}, \theta)$. We observe that ${\mathcal P}^*(M^{2n})$ is stable under the action of the operators $d_k^+, (d _k)^*_{\omega}, d_k ^-$.
\begin{definition}\label{newcomg} Assume that $0\le q \le n-1$.\\
The {\it $k$-plus-primitive q-th cohomology group} of $(M^{2n}, {\omega}, \theta)$ is defined by
\begin{equation}
H^{q} ({\mathcal P}^*(M^{2n}), d_k ^+): = \frac{\ker d_{k} ^+ :{\mathcal P}^{q} (M^{2n}) \to {\mathcal P} ^{q+1}(M^{2n})}{ d_{k}^+ ({\mathcal P}^{q-1} (M^{2n}))}.\label{vy3}
\end{equation}
The {\it $k$-primitive q-th-cohomology group} of $(M^{2n}, {\omega}, \theta)$ is defined by
\begin{equation}
H^{q} ({\mathcal P}^*(M^{2n}), (d_k)^*_{\omega}) : = \frac{\ker (d_k)_{\omega}^*: {\mathcal P}^{q} (M^{2n})\to {\mathcal P} ^{q-1}(M^{2n})}{ (d_{k+1})_{\omega}^* ({\mathcal P}^{q+1}(M^{2n}))}.\label{vy3n}
\end{equation}
The {\it $k$-minus-primitive q-th cohomology group} of $(M^{2n}, {\omega}, \theta)$ is defined by
\begin{equation}
H^{q}({\mathcal P}^*(M^{2n}), d _k ^-):= \frac{\ker d_k^- : {\mathcal P}^q(M^{2n})\to {\mathcal P}^{q-1}(M^{2n})}{d_{k+1}^- ( {\mathcal P}^{q+1}(M^{2n}))}.\label{vy4}
\end{equation}
\end{definition}
Now we show few simple properties of the associated cohomology groups of a l.c.s. manifolds.
Note that the formula (\ref{plus1b}) below has been proved in
\cite[Proposition 3.15]{TY2009} for compact symplectic manifolds $(M^{2n},{\omega})$.
\begin{proposition}\label{plus1} Assume that $(M^{2n}, {\omega}, \theta)$ is a l.c.s. manifold, $n \ge 2$.
1. Suppose that $[(k-1)\theta]\not = 0\in H^1 (M^{2n}, {\mathbb R})$. Then
\begin{equation}
H^{1}({\mathcal P}^*(M^{2n}), d_k ^+) = H^1 ({\Omega}^*(M^{2n}), d_{k}). \label{plus1a}
\end{equation}
2. Suppose that $[(k-1)\theta] = 0\in H^1 (M^{2n}, {\mathbb R})$. Then
\begin{equation}
H^{1} ({\mathcal P}^*(M^{2n}), d _k ^+) = H^1({\Omega}^*(M^{2n}), d_\theta) \text { if } [{\omega}] \not = 0 \in H^2({\Omega}^*(M^{2n}), d_\theta),\label{plus1b}
\end{equation}
\begin{equation}
H^{1} ({\mathcal P}^*(M^{2n}), d_k ^+) = H^1({\Omega}^*(M^{2n}), d_\theta)\oplus R \text { if } [{\omega}] = 0 \in H^2 ({\Omega}^*(M^{2n}), d_\theta),\label{plus1c}
\end{equation}
where $R$ is the 1-dimensional vector space generated by $\rho \in H^{1} ({\mathcal P} ^*(M^{2n}), d_k ^+)$ with $d _{k} \rho = {\omega}$.
\end{proposition}
\begin{proof} 1. Assume that $0\not=\alpha \in {\mathcal P}^1(M^{2n})$ and $d_k ^+ \alpha = 0$, i.e. $[\alpha] \in H^{1} ({\mathcal P} ^*(M^{2n}), d_k ^+)$. Since $d^+_k \alpha = 0$ we get
$d_k \alpha = L f$, where $f\in C^\infty (M^{2n})$. Assume that $f \not = 0$.
Using $d^2_k \alpha = 0$ we derive $Ld_{k-1}f = 0$, which implies $d_{k-1}f =0 $, since $L$ is injective. The equality $d_{k-1} f = 0$ implies that $d_{k-1}$ is gauge equivalent to $d$. This contradicts
the assumption of Proposition \ref{plus1}.1. Hence $f = 0$. It follows $d _k \alpha = 0$. Using $d ^+_kh = d _k h$ for all $h \in {\mathcal P}^0 (M^{2n})$ we
obtain (\ref{plus1a}) immediately.
2. Now we assume that $[(k-1)\theta] = 0\in H^1 (M^{2n}, {\mathbb R})$, or equivalently, $d_{k-1}$ is gauge equivalent to the canonical connection : $d_{k-1} = e^h d e^{-h} = d - dh\wedge$ for some $h \in C^\infty (M ^{2n})$. In this case, as above, $d _{k}\alpha = Lf$ implies $d_{k-1}f = 0$, and hence
$f = ce^{h}$.
a) Assume that (\ref{plus1b}) holds. If $c \not = 0$, then $[e^h{\omega}] = 0 \in H^2(M^{2n}, d_k)$. Since $(k-1)\theta = -dh$, the deformed differential $ d _\theta - dh\wedge$ is gauge equivalent to $d {_k}$.
It follows that $[{\omega}] = 0 \in H^2 (M^{2n}, d_\theta)$. This contradicts the assumption of (\ref{plus1b}).
Hence $c =0$. In this case we have $d_{k} \alpha = 0$, and therefore $[e^{-h}\alpha] \in H^1(M^{2n}, d_\theta)$. Taking into account $d_k^+ f = d_k f$ for any
$f \in {\mathcal P}^0 (M^{2n}) = {\Omega}^0 (M^{2n})$, we obtain (\ref{plus1b}).
b) Assume that (\ref{plus1c}) holds. Then $e^h{\omega} = d_k \rho$ for some $\rho \in {\Omega}^1 (M^{2n})$.
In this case, $d_k(\alpha) = L f = {\omega} c e ^h $ implies $d_k (\alpha - c\rho) = 0$. We conclude that if $d_k ^+(\alpha) = 0$ then $\alpha = c \rho + \beta$ where
$d_k (\beta) =0$. Clearly $[\beta] \in H^1(M^{2n}, d_k) = H^1(M^{2n}, d_\theta)$.
This proves (\ref{plus1c}) and completes the proof of Proposition \ref{plus1}.
\end{proof}
\begin{proposition}\label{primi} (cf. \cite[Lemma 2.7, part II]{TY2009}) Assume that $0\le k \le n$.
If $\alpha \in {\mathcal P}^k (M^{2n})$, then
\begin{equation}
d^-_r (\alpha ^k) = {(d_r)^*_{\omega} (\alpha ^k)\over n-k+1}. \label{plus13}
\end{equation}
Consequently $H^{k} ({\mathcal P}^*(M^{2n}), d^-_r) = H^{k}({\mathcal P}^*(M^{2n}), (d_r) ^* _{\omega})$.
\end{proposition}
\begin{proof} It suffices to prove (\ref{plus13}) locally. Note that locally $d _\theta = d - df\wedge$. In this case $(d-df\wedge){\omega} = 0$ implies ${\omega} = e^f {\omega}_0$ with $d {\omega}_0 = 0$. Next, we compare $*_{\omega}$ and $*_{{\omega}_0}$, using (\ref{sst}) and the equality $G_{\omega} = e ^{-f}G_{{\omega}_0}$.
$$\beta^k \wedge *_{\omega} \alpha^k = \wedge ^k G_{\omega} (\beta^k, \alpha^k) \wedge \frac {{\omega} ^n} { n !} = \wedge ^k e ^{-kf}G_{{\omega}_0}(\beta^k, \alpha ^k) e^{nf}\frac {{\omega}_0 ^n} { n !}= e^{(n-k)f} \beta^k\wedge*_{{\omega}_0} \alpha^k,$$
where $\beta^k , \alpha ^k \in {\Omega} ^k (M^{2n})$. It follows that
\begin{equation}
* _{\omega}(\alpha ^k) = e ^{(n-k)f}*_{{\omega}_0} (\alpha ^k).\label{laho2}
\end{equation}
Let $\alpha^k \in {\mathcal P}^k (M^{2n})$, $0\le k \le n$. Denote by $(d)^*_{{\omega}_0}$ the symplectic adjoint of $d$ with respect to ${\omega}_0$. The formula (\ref{plus15}) below, which is a partial case of (\ref{plus13}) for symplectic manifold, has been proved in \cite[Lemma 2.7, part II]{TY2009}. (We observe that their operator $d^\Lambda$ differs from our operator $(d)^*_{{\omega}_0}$ by sign (-1).)
\begin{equation}
d ^- (\alpha ^k) = { (d )^*_{{\omega}_0} \alpha ^k \over n-k+1}.\label{plus15}
\end{equation}
Using $d_{n+r-k}\alpha = e^{(n+r-k)f}d(e^{-(n +r-k)f}\alpha)$ we obtain from (\ref{laho2})
$$
(d_r)^*_{\omega} (\alpha^k)= (-1) ^k*_{\omega} d_{n+r-k}*_{\omega} (\alpha ^k) =$$
$$= (-1)^k e^{(n - (2n -k+1))f}*_{{\omega}_0} e^{(n+r-k)f}d(e^{-(n +r-k)f} ( e^{(n-k) f}*_{{\omega}_0} \alpha ^k)=$$
\begin{equation}
= (-1) ^k e ^{(r-1) f}*_{{\omega}_0}d ( e^{ -rf}*_{{\omega}_0} \alpha ^k) =\label{gauge1}
\end{equation}
\begin{equation}
= e ^{-f} ( (d)_{{\omega}_0}^* \alpha ^k + (-1) ^k *_{{\omega}_0}(-r) df \wedge *_{{\omega}_0}\alpha ^k). \label{lahor6}
\end{equation}
Substituting $r = 0$, we derive from (\ref{lahor6})
\begin{equation}
(d)_{{\omega}_0} ^* (\alpha ^k) = e^f (d) ^*_{\omega}\alpha ^k . \label{laho5}
\end{equation}
Next we compare $d_r ^-$ with $d^-$.
\begin{equation}
d _r (\alpha^k) = e^{rf} d (e ^{-rf} \alpha^k) = e^{rf} [d^+ (e ^{-rf} \alpha ^k) + {\omega}_0 \wedge d^- e^{-rf} \alpha ^k].\label{laho3}
\end{equation}
Since the Lefschetz decomposition of ${\Omega}(M^{2n})$ is invariant under conformal transformations we obtain from (\ref{laho3})
\begin{equation}
d_r^- (\alpha ^k) = e^{(r-1)f} d^- (e ^{-rf} \alpha ^k).\label{laho1}
\end{equation}
Combining (\ref{laho1}) with (\ref{plus15}) we conclude that
\begin{equation}
d_r ^- (\alpha ^k)= e^{(r-1)f}\frac{(d)_{{\omega}_0}^* (e^{-rf} \alpha ^k)}{n-k+1}.\label{laho7}
\end{equation}
Taking into account (\ref{laho5}) and (\ref{gauge1}), we derive from (\ref{laho7})
\begin{equation}
d^-_r (\alpha ^k) = e^{rf}\frac{(d)^*_{{\omega}}e^{-rf} \alpha ^k}{ n-k+1} = \frac{(d_r)^*_{\omega} \alpha ^k}{n-k+1} .\label{plus16}
\end{equation}
This proves (\ref{plus13}).
Clearly the second assertion of Proposition \ref{primi} follows from (\ref{plus13}). This completes the proof of Proposition \ref{primi}.
\end{proof}
Let $J$ be a compatible almost complex structure on a l.c.s. manifold $(M^{2n}, {\omega}, \theta)$. The complexified space $T^*_{\mathbb C} (M^{2n}) : =(T ^*(M ^{2n})\otimes {\mathbb C} $ is decomposed into
eigen-subspaces $T^{p,q}(M^{2n})$. Let $\Pi^{p,q}: T_{\mathbb C}^*(M^{2n}) \to T^{p,q} (M^{2n})$ be the projection.
Set
$${\mathcal J} : = \sum _{p,q} (\sqrt{-1})^{p-q} \Pi ^{p,q}.$$
In what follows we want to apply the Hodge theory for compact l.c.s. manifolds $(M^{2n}, {\omega}, \theta)$ provided with a compatible metric $g$. First we derive a formula for the formal adjoint $(d ^+ _l)^*$ of $d ^+ _l: {\mathcal P}^*(M^{2n}): = {\mathcal P} ^* (M^{2n}) \to {\mathcal P}^*(M^{2n})\subset {\Omega}^*(M^{2n})$. For any operator $D$ acting on a subbundle $E \subset {\Omega}^* (M^{2n})$ we denote by $(D)^*$ the formal adjoint of $D$.
\begin{lemma}\label{formal1} For any $\alpha \in {\mathcal P}^* (M^{2n})$ we have
\begin{eqnarray}
(d^+ _l) ^* (\alpha) = - *_g (d _{-l}) *_g(\alpha). \label{formal3}
\end{eqnarray}
\end{lemma}
\begin{proof} First, we want to compute the formal adjoint $(d_l) ^* $ of $d _l = d + l \theta \wedge : {\Omega} ^* (M^{2n}) \to {\Omega} ^* (M^{2n})$. It is known that \cite[\S 5.1.2]{Voisin2007}
\begin{equation}
(d)^* = - * _g d *_g.\label{voisin521}
\end{equation}
Since $\theta \wedge$ is the symbol of $d$ we derive from (\ref{voisin521})
\begin{equation}
(l\theta \wedge ) ^* = *_g l\theta \wedge *_g. \label{formal7}
\end{equation}
It follows from (\ref{voisin521}) and (\ref{formal7})
\begin{equation}
(d _l )^*= -*_gd _{-l}*_g.\label{formal4}
\end{equation}
Using (\ref{lel}) we get
for $\alpha \in {\mathcal P} ^* (M^{2n})$
\begin{equation}
(Ld_l ^-) ^*( \alpha) = (d_l ^-)^* \Lambda (\alpha) = 0.\label{formal9}
\end{equation}
It follows from (\ref{formal4}) and (\ref{formal9}) that for $\alpha \in {\mathcal P} ^* (M^{2n})$
\begin{equation}
(d_l^+) ^*(\alpha) = -*_g (d_{-l}) *_g(\alpha) . \label{formal10}
\end{equation}
This proves (\ref{formal3}), which is consistent with \cite[(3.2), part II]{TY2009}, if $(M^{2n}, {\omega})$ is a symplectic manifold.
\end{proof}
\begin{lemma}\label{dpm} (cf. \cite[Lemma 3.4, part II]{TY2009}) Let $J$ be a compatible almost complex structure on a l.c.s. manifold $(M^{2n}, {\omega}, \theta)$, $g$ the associated compatible metric and $*_g$ the Hodge star operator with respect to $g$.
Then for $\alpha ^k, \alpha ^{k-1} \in {\mathcal P}^* (M^{2n})$, $0\le k \le n$, we have
\begin{eqnarray}
{\mathcal J} (d^+_l) ^* {\mathcal J}^{-1}(\alpha^k) = (n-k+1) d^- _{-l+k -n}(\alpha^k),\label{conj2}\\
{\mathcal J} d ^+_l {\mathcal J}^{-1} (\alpha ^{k-1}) = (n-k+1) (d^- _ {-l+k-n}) ^*(\alpha^{k-1}). \label{conj1}
\end{eqnarray}
\end{lemma}
\begin{proof} Using \cite[Theorem 2.4]{Brylinski1988}, see also Lemma \ref{hk}, we get easily
\begin{equation}
{\mathcal J} = *_g *_{\omega}. \label{conj3}
\end{equation}
By (\ref{voisin5.5}) we derive from (\ref{conj3})
\begin{equation}
{\mathcal J} ^{-1} (\alpha ^k) = *_{\omega} *_g (-1) ^k(\alpha^k).\label{conj3s}
\end{equation}
Combining (\ref{conj3s}) with (\ref{conj3}), (\ref{formal3}) and applying (\ref{defnbls}), (\ref{voisin5.5}) again we obtain
\begin{eqnarray}
{\mathcal J} (d^+ _l)^*_g {\mathcal J} ^{-1} (\alpha ^k) = (-1) ^{k+1} *_g *_{\omega} *_g d _{-l} *_g *_{\omega} *_g(\alpha^k)=\nonumber\\
= (-1)^{k+1}*_g ^2 ( d_{-l+k -n}) ^*_{\omega} (\alpha^k) = (d_{-l +k -n})^*_{\omega} (\alpha^k),\label{conj4}
\end{eqnarray}
since $*_{\omega} *_g = *_g *_{\omega}$.
Using (\ref{plus13}) we derive (\ref{conj1}) immediately from (\ref{conj4}). Clearly (\ref{conj1}) follows from (\ref{conj2}), since they are adjoint. This completes the proof of Lemma \ref{dpm}.
\end{proof}
The following Proposition is a generalization of \cite[Proposition 3.5, part II]{TY2009}.
\begin{proposition}\label{dpm1} Let $(M^{2n}, {\omega}, \theta)$ be a compact l.c.s manifold. Then there is $H ^{k} ({\mathcal P} ^*(M^{2n}), d^+_l) = H^{k} ({\mathcal P} ^*(M^{2n}),(d_{-l+k -n})^*_{\omega} )$ for all $ l \in \Z$ and $0\le k \le n-1$.
\end{proposition}
\begin{proof} First we note that all the operators $d ^+_l$, $d^-_l$ and $(d_l)^* _{\omega}$ restricted to the space ${\mathcal P}^*(M^{2n})$ are elliptic operators. This observation is a consequence of
the theorem by Bouche who proved that the complex of coeffective forms on a symplectic manifold $M^{2n}$ is elliptic in dimension greater than $n$ \cite{Bouche1990}. Indeed, the complex $({\mathcal P}^*(M^{2n}), (d)_{\omega} ^*)$ is dual to the complex
of coeffective forms, see also Remark \ref{his1}.1 below. Thus $(d_l) ^*_{\omega}$ acting on ${\mathcal P} ^*(M^{2n})$ is an elliptic operator, since $(d_l)^*_{\omega}$ has the same symbol as $(d)_{\omega}^*$. Taking (\ref{plus13}), (\ref{conj1}) and (\ref{conj2}) into account we prove the ellipticity of $d^-_l$ and $d^+_l$ acting
on ${\mathcal P}^*(M^{2n})$. In \cite[Proposition 2.8 part II]{TY2009} the authors give another proof of the ellipticity of these operators.
Now Proposition \ref{dpm1} follows easily from Lemma \ref{dpm} using the Hodge theory.
\end{proof}
\begin{corollary}\label{lich0} Assume that $(M^{2n}, {\omega}, \theta )$ is a connected compact l.c.s. manifold. Then $H^{0} ({\mathcal P}^*(M^{2n}), d^+_k) = 0 $ if $d_k$ is not gauge equivalent to the
canonical differential $d = d_0$, ore equivalently $[k\theta] \not = 0\in H ^1 (M^{2n}, {\mathbb R})$.
If otherwise, then $H^{0} ({\mathcal P}^*(M^{2n}), d^+_k) = H^{0} ({\mathcal P} ^* (M^{2n}), (d_k)^*_{\omega}) = H^0 ({\mathcal P} ^*(M^{2n}), d_k ^-)= {\mathbb R}$.
\end{corollary}
\begin{proof} Note that ${\mathcal P}^0 (M^{2n}) = {\Omega} ^0 (M^{2n})$ and $d ^+_k = d_k$, which implies the first assertion of Proposition
\ref{lich0} immediately. The second assertion of Proposition \ref{lich0} follows from Proposition \ref{primi} and Proposition \ref{dpm1}, taking into account the equalities $H^{0} ({\mathcal P} ^*(M^{2n}), d) = H^0(M^{2n}, {\mathbb R} ) = {\mathbb R}$.
\end{proof}
\begin{remark} \label{his1} 1. Let $(M^{2n}, {\omega}, \theta)$ be a l.c.s. manifold. The symplectic star operator $*_{\omega}$ provides an isomorphism between the space ${\mathcal P}^*M^{2n}$ of primitive forms and the space $\Cc ^* M^{2n}$ of
coeffective forms. Thus $H^{*}({\mathcal P}^* (M^{2n})) $ is isomorphic to $H^*(\Cc ^* M^{2n}, d_\theta)$. The latter cohomology groups for symplectic manifolds
have been introduced by Bouche \cite{Bouche1990}. A variant of the effective cohomology groups for contact manifolds has been introduced (and computed) by Lychagin \cite{Lychagin1979} already in 1979. Later, a modification of this complex for contact manifolds has been considered by Rumin independently \cite{Rumin1994}. Chinea, Marrero and Leo generalized the construction of effective cohomology groups for Jacobi manifolds \cite{CML1998}.
2. Note that the groups $H^{q} ({\mathcal P}^*(M^{2n}),d^+_k)$ have
the following simple interpretation. We consider the differential ideal $ L ({\Omega} ^* (M^{2n})) \subset {\Omega}^*(M^{2n})$. The quotient ${\Omega}^* (M^{2n})/ L({\Omega}^*(M^{2n}))$ is isomorphic to the space ${\mathcal P}^*(M^{2n})$,
and the differential $d_k$ induces the differential $d_k ^+$ on the quotient complex.
3. The plus-primitive cohomology groups and the minus-primitive cohomology groups for symplectic manifolds
have been introduced by Tseng and Yau \cite[part II]{TY2009}.
4. Below we shall show a deeper relation between these new cohomology groups and the twisted cohomology groups $H^*(M^{2n}, d_k)$ using the spectral sequence introduced in the next section.
5. Let $(M^{2n+1}, \alpha)$ be a contact manifold. Then its symplectization $M^{2n+2}:= M^{2n+1} \times {\mathbb R}$ is supplied with a symplectic form ${\omega}(x, t) = \exp ^t (d\alpha + dt \wedge \alpha) = \tilde \alpha$.
Denote by $i : M^{2n+1} \to M^{2n+2}$ the embedding $x \mapsto (x, 0)$. We observe that the restriction of the
filtration on $(M^{2n +2}, d \tilde \alpha)$ to $i (M^{2n+1})$ coincides with the filtration introduced
by Lychagin \cite{Lychagin1979}.
6. Note that any symplectic manifold $(M^{2n}, {\omega})$ is a Poisson manifold equipped with the Poisson bivector $G_{\omega}$. Associated with a Poisson bivector $\Lambda$ on a Poisson manifold
$M$ there are two differential complexes. The first one is the complex of multivector fields on $M$ equipped with the differential $\Lambda$ acting via the Schouten bracket. It has been introduced by Lichnerowicz and the associated cohomology group is called
the Lichnerowicz-Poisson cohomology of $M$ \cite{Lichnerowicz1977}, \cite{FIL1996}. The second differential complex is the complex of differential forms on $M$ equipped with the differential $\delta = [i(\Lambda), d]$ where
$i(\Lambda)$ is the contraction with $\Lambda$. This complex has been introduced by Kozsul in \cite{Koszul1985} and it is called the canonical Poisson homology of $M$ \cite{Brylinski1988}.
If $M^{2n}$ is symplectic then $G_{\omega}\in End (T^*M^{2n}, TM^{2n})$ induces an isomorphism between the de Rham cohomology and Lichnerowicz-Poisson cohomology \cite[Theorem 6.1]{FIL1996}, and the symplectic star operator provides an isomorphism between the de Rham cohomology and the canonical Poisson homology \cite{Brylinski1988}. In \cite{FIL1996} the authors consider the coeffective Lichnerowicz-Poisson cohomology groups on a Poisson manifold, which are isomorphic to the coeffective symplectic groups introduced by Bouche \cite{Bouche1990} if the Poisson structure is symplectic.
\end{remark}
\section{Spectral sequences on a l.c.s. manifold}
In this section, for any integer $k$, we construct a spectral sequence associated with the filtered complex $(F^*K^*, d_k)$ on a l.c.s. manifold $(M^{2m}, {\omega}, \theta)$. We compare the $E_1$-term of this spectral sequence with the primitive cohomology group $H^{*,+}({\mathcal P}^*(M^{2n}), d_k)$ introduced in the previous section (Lemma \ref{fk1}).
We show the existence of a long exact sequence connecting the $E_1$-term of this spectral sequence with the Lichnerowicz-Novikov cohomology
groups $H^* ({\Omega}^*(M^{2n}), d_{k+p})$ for appropriate $p$, which will be denoted by $H^* _{k+p}(M^{2n})$ (Theorem \ref{ex1}, Proposition \ref{prol1}). We prove that the $E_1$-term of our spectral sequence
is a conformal invariant of $(M^{2n},{\omega}, \theta)$, moreover, the $E_1$-terms associated with $(M^{2n}, {\omega}, \theta)$
and $(M^{2n}, {\omega}', \theta)$ are isomorphic, if ${\omega}'= {\omega} + d_{\theta} \tau$ (Theorem \ref{conf}).
In Proposition \ref{clo2} of the previous section we proved that $(F^*K^*, d_k)$ is a filtered complex.
Let us study the spectral sequence $\{ E_{k,r}^{p,q}, d_{k,r}: E^{p,q}_{k,r} \to
E^{p+r,q-r +1}_{k,r}\}$ of this filtered complex, first introduced by Di Pietro and Vinogradov for symplectic manifolds in \cite{PV2006}. We refer the reader to \cite{McCleary2001}, \cite{GH1978} for an introduction
into the theory of spectral sequences associated with a filtration.
The initial term $E^{p,q}_{k,0}$ of this spectral sequence is defined as follows,
\begin{equation}
E^{p,q}_{k,0} = F^p K^{p+q}/F^{p+1} K^{p+q}.\label{e00}
\end{equation}
Using the induced Lefschetz decomposition (\ref{fil1}), (\ref{fil1a}),
taking into account the injectivity of the map $L ^p : {\Omega} ^{q-p}(M^{2n}) \to {\Omega}^{q+p} (M^{2n})$, we get for all $k \in \Z$
\begin{equation}
E^{p,q}_{k,0} \cong \Ll^{p, q-p} \cong \Ll ^{0, q-p} \text{ if } n \ge q \ge p \ge 0,
\label{e01}
\end{equation}
\begin{equation}
E^{p,q}_{k,0} = 0 \text{ otherwise }.
\label{e02}
\end{equation}
\medskip
Since $E^{p,q}_{k,l}$ is a quotient of $E^{p,q}_{k, l-1}$, in view of (\ref{e02}), any term $E^{p,q}_{k,l}$ written below, if without explicit condition on $p$ and $ q$, is always under the assumption $ 0\le p \le q \le n$.
Now let us go to the next term $E^{p,q}_{k,1}$ of our spectral sequence.
Recall that $d_{k, 0} : E^{p,q}_{k,0} \to E^{p, q+1}_{k,0} $ is obtained from the differential $d_{k}$ by passing to the quotient:
\begin{equation}
\xymatrix{E^{p,q}_{k,0} \ar[r]^{d_{k,0}}\ar@{=}[d] & E^{p, q+1}_{k,0}\ar@{=}[d] \\
F^p K^{p+q}/F^{p+1}K^{p+q}\ar[r] & F^p K^{p+q+1}/F^{p+1}K^{p+q+1} &
}
\label{e0d}
\end{equation}
Let us write $d_{k,0}$ explicitly using (\ref{e01}) and (\ref{e02}).
Since $d_k L^p = L^p d_{k-p}$, using (\ref{e00}), (\ref{e01}), (\ref{e02}) and (\ref{hoya}) we have for any $\alpha \in E^{p,q}_{k,0}$ with $n \ge q \ge p \ge 0$
\begin{equation}
d_{k,0} (\alpha) =[L^p (d_{k-p} ^+ + L d_{k-p} ^-)( \tilde \alpha)] = [L^p d_{k-p} ^ + ( \tilde \alpha)]\in E^{p, q+1}_{k,0}, \label{vy2}
\end{equation}
where $\tilde \alpha \in \Ll^{0, q-p}$ is a representative of $\alpha \in E^{p,q}_{k,0}$ by (\ref{e01}).
Since $L^p : \Ll ^{0, q-p} \to \Ll^{p, q-p}$ is an isomorphism, if $0\le p \le q \le n$ by (\ref{e01}), we rewrite (\ref{vy2}) as follows
\begin{eqnarray}
d_{k,0}: E^{p,q}_{k,l}= \Ll ^{0, q-p} \to E^{p, q+1}_{k, 0} = \Ll ^{0, q+1 -p}_{ k, 0}, \, \tilde \alpha \mapsto d^+_{k-p}\tilde \alpha, \label{e12} \\
\text{ if } 0\le p \le q \le n. \nonumber
\end{eqnarray}
\begin{lemma}\label{fk1} The term $E^{*,*}_{k,1}$ of the spectral sequence $\{ E^{p,q}_{k,r}, d_{k,r}\}$ is determined by the following relations
\begin{equation}
E^{p,q}_{k,1} = H^{q-p} ({\mathcal P}^*(M^{2n}), d^+_{k-p}) \text { if } 0 \le p \le q \le n-1,
\label{e13}
\end{equation}
\begin{equation}
E^{p,n}_{k,1} = { {\mathcal P}^{n-p}(M^{2n}) \over d_{k-p}^+ ({\mathcal P}^{n-p-1}(M^{2n}))}, \text { if } 0 \le p \le n,\label{e13a}
\end{equation}
\begin{equation}
E^{p,q}_{k,1} = 0 \text { otherwise }. \label{e13b}
\end{equation}
\end{lemma}
\begin{proof} Clearly (\ref{e13}) is a consequence of (\ref{e12}).
Next using (\ref{vy2}) and the identity $L ^p( {\mathcal P}^{n -p +1 } (M^{2n})) = 0$, we obtain $d_{k, 0} (E^{p, n} _{k,0} ) = 0$. Hence
\begin{equation}
E^{p,n}_{k,1} = { L^p({\mathcal P}^{n-p}(M^{2n})) \over L^p ( d_{k-p}^+ ({\mathcal P}^{n-p-1}(M^{2n})))}. \label{e13aa}
\end{equation}
Since $L^p : {\Omega} ^{n-p} (M^{2n}) \to {\Omega} ^{n+p} (M^{2n})$ is injective, (\ref{e13aa}) implies (\ref{e13a}).
The last relation (\ref{e13b}) in Lemma \ref{fk1} follows from (\ref{e02}). This completes the proof of Lemma \ref{fk1}.
\end{proof}
Next we define the following diagram of chain complexes
$$
\xymatrix{ {\Omega} ^{q-p-1}(M^{2n})\ar[r] ^ {L} \ar[d] ^{d_{l-1}} & {\Omega} ^{q-p+1}(M^{2n}) \ar[r]^{\Pi L^p}\ar[d]^{d_{l} } & E^{p, q+1}_{l+p,0}\ar[d]_{d_{l+p,0}}\ar[r]&\\
{\Omega} ^{q-p}(M^{2n}) \ar[r] ^L\ar[d]^{d_{l-1}} & {\Omega} ^{q-p+2}(M^{2n}) \ar[r]^ {\Pi L^p}\ar[d]^{d_{l}} &E^{p, q+2}_{l+p,0} \ar[d]_{d_{l+p,0}}\ar[r] & \\
{\Omega} ^{q-p+1}(M^{2n})\ar[r] ^L & {\Omega} ^{q-p+3}(M^{2n}) \ar[r]^{\Pi L^p} & E^{p, q+3}_{l+p,0} \ar[r] &
}
$$
Here the map $\Pi L^p$ associates with each element $\beta \in {\Omega}^{q-p+1}(M^{2n})$ the element $[L ^p\beta]\in
E^{p, q+1}_{l+p,0}$.
Recall that $d_{l} \circ L = L \circ d_{l-1}$. Thus the upper part of the above diagram is commutative.
The lower part of the diagram is also commutative, since
$$ d_{l+p} L^ p = L^p d_{l} .$$
Summarizing we have the following sequence of chain complexes
\begin{equation}
0\to ({\Omega}^{q-(p+1)}(M^{2n}), d_{l-1}) \stackrel{L}{\to} ({\Omega}^{q+1 -p}(M^{2n}), d_{l}) \stackrel{\Pi L^p}{\to} (E_{l+p,0}^{p, q+1}, d_{l+p,0}) \to 0.\label{exac3}
\end{equation}
Set ${\Omega} ^{-1} (M^{2n}): = 0$.
\begin{lemma}\label{lexac3} The sequence (\ref{exac3}) of chain complexes is exact for $0\le p \le q \le n-1$.
\end{lemma}
\begin{proof}
For $0\le p \le q \le n$ the operator $L: {\Omega} ^{q-p-1} (M^{2n}) \to {\Omega} ^{q-p+1}(M^{2n})$ is injective by Lemma \ref{inj1}, so the sequence
(\ref{exac3}) is exact at ${\Omega} ^{q -(p+1)}(M^{2n})$. The exactness at ${\Omega}^{q+1-p}(M^{2n})$ follows easily from Corollary \ref{inj1}, taking into account (\ref{e01}). The exactness at $E^{p,q+1}_{l+p,0}$ follows directly from the definition (\ref{e00}) of $E^{p,q}_{l+p,0}$.
\end{proof}
As a consequence of Lemma \ref{lexac3}, using the general theory of homological algebra, see e.g. \cite[Chapter 1, \S 6]{GM1988}, we get immediately
\begin{theorem}\label{ex1} The following long sequence is exact for $0\le p \le q \le n-1$
\begin{equation}
\cdots \to H^{q-p}_l(M^{2n}) \stackrel{ \bar L^p}{\to} E^{p,q}_{l+p,1}\stackrel{\delta_{p,q}}{\to} H^{q-(p+1)}_{l-1}(M^{2n})\stackrel{\bar L} {\to} H^{q+1-p}_{l}(M^{2n}) \stackrel{\bar L^p} {\to } E^{p, q+1}_{ l+p,1} \to \cdots \label{ex11}
\end{equation}
\end{theorem}
\begin{remark}\label{rpv} 1. Theorem \ref{ex1} is a generalization of \cite[Theorem 1]{PV2006} stated for symplectic manifolds.
2. Let us write the connecting homomorphism $\delta_{p,q}$ explicitly. If $\alpha \in H^{q-p} ({\mathcal P}^*(M^{2n}), d_l^+) =
E^{p,q}_{l+p,1}$, see (\ref{e13}), then $d_l\tilde \alpha = L \eta $ for any representative
$\tilde \alpha \in {\mathcal P}^{q-p}(M^{2n})$ of $\alpha$. By the definition of connecting homomorphism, see also \cite[\S 3]{PV2006}, $\delta_{p,q} (\alpha) = [\eta ] \in H^{q-(p+1)}_{l-1} (M^{2n})$. Since $d^+_l\tilde \alpha = 0$, $d_l \tilde \alpha = Ld_l ^- \tilde \alpha$, so $\eta = d^-_l \tilde \alpha$. Thus we get
\begin{equation}
\delta_{p,q} \alpha = [d_{l}^- \tilde \alpha] \in H^{q-(p+1)}_{l-1} (M^{2n}).\label{delta1}
\end{equation}
3. Let us write the operator $\bar L^p$ explicitly. Assume that $[\beta] \in H^{q-p} _l (M^{2n})$, $\beta \in {\Omega}^{q-p} (M^{2n})$. Set
$$\beta _{pr}: = \Pi_{pr} \beta$$
- the primitive component of $\beta$ in the first Lefschetz decomposition. Since $d_{l}\beta = 0$, we have $d_{l} ^ + \beta _{pr} = 0$.
Thus the image $\bar L^p [\beta] = [L^p \beta] \in E^{p,q}_{l+p, 1}=H^{q-p} ({\mathcal P}^*(M^{2n}), d_l^+)$ has a representative $\beta _{pr}\in {\mathcal P}^{q-p}(M^{2n})$ with
$d_l^+\beta _{pr} = 0$. Summarizing we have
\begin{equation}
\bar L^p [\beta] = [\beta_{pr}] \in H^{q-p} ({\mathcal P}^*(M^{2n}), d^+_l) = E^{p,q}_{l+p, 1}. \label{lpex}
\end{equation}
\end{remark}
4. Substituting for $0\le p= q \le n-1$ in the long exact sequence (\ref{ex11}), we get the left end of (\ref{ex11})
\begin{equation}
0 \to H^0_l (M^{2n}) \to E^{p,p}_{l+p, 1} \to 0 \to H^1 _l (M^{2n}) \to \cdots . \label{dege1}
\end{equation}
From (\ref{e13}) and (\ref{dege1}) we get for $0\le p \le n-1$
\begin{equation}
E^{p, p}_{l+p, 1}= H^{0,+} ({\mathcal P}^*(M^{2n})) = H ^0_l (M^{2n}) .\label{0pl}
\end{equation}
\medskip
Let us prolong the exact sequence (\ref{ex11}) for $q=n$, using the ideas in \cite[\S III]{PV2006}.
For $0 \le k \le n$ we set
$$C^k_l : = {\ker d_l ^- \cap {\Omega} ^k (M^{2n})\over d_l ( {\Omega} ^{k-1} (M^{2n}))}.$$
\begin{proposition}\label{prol1} The long exact sequence (\ref{ex11}) can be extended as follows
\begin{equation}
E^{p, n-1}_{l+p, 1}\stackrel{\delta_{p,n-1}}{ \to} H^{n - (p+2)}_{l-1}(M^{2n}) \stackrel{[L]}{\to} C^{n-p}_{l}\stackrel{[ L^p] }{\to} E^{p, n}_{l+p,1} \stackrel{\delta_{p,n}}{\to} C^{n - (p+1)}_{l-1} \stackrel{[ L^{p+1}] }{\to} H^{n +p +1}_{l+p} (M^{2n}) \to 0,\label{prol1a}
\end{equation}
where $0\le p \le n-1$ and the operators $[L], [L^p]$ and $[L^{p+1}]$ will be defined in the proof below.
\end{proposition}
\begin{proof} First we define $[L]$ and prove the exactness at $H^{n-(p+2)}_{l-1} ( M^{2n})$. Denote by $\Pi: H^{n-p}_{l} (M^{2n}) \to C^{n-p}_{l}$ the natural embedding of the quotient spaces
$$\frac{\ker d_l \cap {\Omega} ^{n -p} (M^{2n})}{d_l ({\Omega} ^{n-p-1} (M^{2n}))} \to \frac{\ker d_l^- \cap {\Omega} ^{n -p} (M^{2n})}{d_l ({\Omega} ^{n-p-1} (M^{2n}))}.$$
Set $[L]: = \Pi \circ \bar L$, where $\bar L:H^{n - (p+2)}_{l-1}(M^{2n}) \to H^{n-p}_{l}(M^{2n})$ is induced by $L$. By Theorem \ref{ex1} we
have $Im \, \delta _{p, n-1} = \ker \bar L$.
Since $\Pi$ is an embedding, the last equality implies $\ker [ L ]= \ker \bar L$. This proves the required exactness at $H^{n-(p+2)}_{l-1} ( M^{2n})$.
Now we define $[L^p]$ and show the exactness at $C^{n-p}_{l}$. Assume that $\alpha = \alpha _{pr} + L \tilde \alpha \in {\Omega} ^{n-p} (M^{2n})$ is a representative of $[\alpha]\in C^{n-p}_{l}$,
i.e. $ d_{l}^- (\alpha ) = 0$, or equivalently $d_l \alpha = d_l ^+ \alpha$.
We set
\begin{equation}
[L^p] (\alpha): = [\alpha_{pr}]\in \frac{{\mathcal P}^{n-p}(M^{2n})}{ d _{l}^+ ({\mathcal P}^{n-p-1}( M^{2n}))} = E^{p,n} _{l+p, 1} .\label{new1a}
\end{equation}
Clearly the map $[L^p]$ is well-defined, since $\Pi_{pr}d_l (\gamma) = d_l ^+ \Pi_{pr}(\gamma)$.
Now assume that $[\alpha] \in \ker [L^p]$, so by (\ref{new1})
\begin{equation}
\Pi_{pr}\alpha = d^+_l \gamma \text { for some }\gamma \in {\mathcal P} ^{n-p-1} (M^{2n}).\label{new11}
\end{equation}
Using the property $d_l \alpha = d_l ^+ \alpha$ we obtain from (\ref{new11}) $d_l \alpha = 0$.
Now we write
\begin{equation}
\alpha = \alpha _{pr} + L\tilde \alpha = d_l ^+ \gamma + L\tilde \alpha = d_l \gamma + L \beta,\label{new2}
\end{equation}
where $\beta = d_l^- \gamma + \tilde \alpha$.
Since $d_l \alpha = 0$, using (\ref{new2}) we get $d_l L\beta = L d_{l-1} \beta = 0$. Applying Lemma \ref{inj1} to $d_{l-1}\beta \in {\Omega}^{n-p-1}(M^{2n})$, we obtain $d_{l-1} \beta = 0$. This implies $[\alpha] = [L] ([\beta]) \in {\rm Im}\, [L]$, and the required exactness.
Next we define the operator $\delta_{p,n}:E^{p, n}_{l+p,1} \to C^{n - (p+1)}_{l-1} $ as follows
\begin{equation}
\delta_{p,n}(\alpha ) := [d_l ^- \tilde \alpha] \in C^{n - (p+1)}_{l-1},\label{deltac}
\end{equation}
where $\tilde \alpha \in {\mathcal P}^{n-p}( M^{2n})$ is a representative of $\alpha \in E^{p,n} _{l+p,1}= {\mathcal P}^{n-p}(M^{2n})/ d _{l}^+ ({\mathcal P}^{n-p-1} (M^{2n}))$.
Clearly $\delta_{p,n} (\alpha) \in C^{n-p-1}_{l-1}$, since by (\ref{com1}) $d_{l-1} (d_l^-\tilde \alpha ) = d^+_{l-1} d^-_{l } \tilde \alpha$.
The map $\delta_{p,n}$ is well-defined, since for any $\gamma \in {\mathcal P}^{n-p-1}(M^{2n})$ using (\ref{com1}) and (\ref{com1a}) we get
\begin{equation}
[d_l ^- d_l^+ \gamma] = -[d_{l-1}^+ d_l ^- \gamma] = -[d_{l-1} (d_l^- \gamma)] = 0 \in C^{n-p-1}_{l-1}.
\end{equation}
Now assume that $\alpha \in \ker \delta _{p,n}$ and $\tilde \alpha \in {\mathcal P}^{n-p}(M^{2n})$ is its representative. By (\ref{deltac}) $d_{l} ^- \tilde \alpha = d_{l-1} \beta$ for some
$\beta \in {\Omega} ^{n-p-2} (M^{2n})$. It follows
\begin{equation}
d_{l} ^+ \tilde \alpha = d_l \tilde \alpha - L d_{l-1} \beta = d_l ( \tilde \alpha - L\beta).\label{deltac1}
\end{equation}
Since $\tilde \alpha$ is primitive, and $d^+_l (\Ll ^{s, r}) \subset \Ll ^{s, r+1}$, we get
\begin{equation}
d_{l}^+ \tilde \alpha = d_l ^+ (\tilde \alpha - L \beta).\label{deltac2}
\end{equation}
Clearly, (\ref{deltac1}) and (\ref{deltac2}) imply $[\tilde \alpha - L \beta] \in C^{n-p}_l$, and by
(\ref{new1}) $\alpha = [L^p] ([\tilde \alpha - L \beta])$. This yields the exactness at $E^{p, n}_{l+p,1}$.
Let us define $[L^{p+1}]$ and show the exactness at $C^{n-(p+1)}_{l-1}$. For $\alpha \in C^{n-p-1}_{l-1}$ we set
\begin{equation}
[L^{p+1}](\alpha): = [L ^{p+1} \tilde \alpha] \in H^{n+p+1}_{l+p}(M^{2n}),
\end{equation}
where $\tilde \alpha \in {\Omega} ^{n-p-1}(M^{2n})$ is a representative of $\alpha$. Note that $d_{l+p} L^{p+1} \tilde \alpha = L^{p+1}d_{l-1}\tilde \alpha = 0$, so $[L^{p+1}] (\alpha) \in H^{n+p+1}_{l+p}(M^{2n})$. The same formula shows that our map $[L^{p+1}]$ does not depend on the choice
of a representative $\tilde \alpha$ of $\alpha$.
Now assume that $\alpha \in \ker [L^{p+1}]$. Then $L^{p+1}\tilde \alpha = d_{l+p} \beta$
for some $\beta \in {\Omega} ^{n+p} (M^{2n})$. Using the Lefschetz decomposition for $\beta$ we write
$$\beta = L^p(\beta _{0} + \sum_ {k =1} ^{[{n+p\over 2}]}L^k \beta_{k}), \, \beta_i \in {\mathcal P}^*(M^{2n}).$$
It follows that
\begin{equation}
L^{p+1}\tilde\alpha = d_{l+p}\beta = L^p (d^+_{l} \beta_{0} + L (d_{l} ^- \beta_{0} + d_{l-1}( \beta_{1} + L\beta _{2} + \cdots))).\label{deltac3}
\end{equation}
By Corollary \ref{inj1}.1
$L^{p+1}: {\Omega} ^{n-p-1}(M^{2n}) \to {\Omega} ^{n+p+1}(M^{2n})$ is an isomorphism, hence we get from (\ref{deltac3})
\begin{equation}
\tilde\alpha = d_{l} ^- \beta_{p} + d_{l-1} (\beta_{p+1} + L\beta_{p+1} + \cdots ). \label{deltac4}
\end{equation}
Combining (\ref{deltac4}) with (\ref{deltac}) we get $\alpha \in {\rm Im}\, \delta_{p,n}$. This proves the required exactness.
Finally we show that $[L^{p+1}]$ is surjective. Assume that $ \beta \in {\Omega} ^{n+p+1}( M^{2n})$ is a representative of
$[\beta] \in H^{n+p+1}_{l+p+1} (M^{2n})$. Using the Lefschetz decomposition we write $\beta = L^{p+1} (\tilde \beta)$, $\tilde \beta \in {\Omega} ^{n-p-1}(M^{2n})$.
Note that $L^{p+2} d_l ^-\tilde \beta = L d_{l+p} ^- \beta = 0$, since $d_{l+p} \beta = 0$. Since $L^{p+2} : {\Omega} ^{n-p-2}(M^{2n}) \to {\Omega} ^{n+p +2} (M^{2n})$ is an isomorphism, we get $d_l ^- \tilde \beta = 0$. Hence $[\tilde \beta ] \in C^{n-p-1}_l$ and $[\beta ] =[L^{p+1}]( [\tilde \beta])$.
This completes the proof of Proposition \ref{prol1}.
\end{proof}
\begin{theorem}\label{conf} (cf. \cite[Osservazione 18]{Pietro2006}) The spectral sequences $E^{p,q}_{k,r}$ on $(M^{2n}, {\omega}, \theta)$ and
on $(M^{2n}, {\omega} ', \theta')$ are isomorphic, if ${\omega}$ and ${\omega}'$ are conformal equivalent. Furthermore, the $E^{*,*}_{k,1}$-terms of the spectral sequences on $(M, {\omega}, \theta)$ and $(M, {\omega}', \theta')$ are isomorphic, if ${\omega} ' = {\omega} + d_{\theta} \rho$ for some $\rho \in {\Omega} ^1 (M^{2n})$.
\end{theorem}
\begin{proof} If ${\omega}' = \pm e^f {\omega}$, then $d_\theta{\omega} ' = df \wedge {\omega}'$. Hence
\begin{equation}
(d_\theta - df\wedge){\omega}' = 0.\label{confn1}
\end{equation}
Since $L : {\Omega} ^1 (M ^{2n}) \to {\Omega} ^3 (M^{2n})$ is injective, (\ref{confn1}) implies that
\begin{equation}
d_{\theta '} = d_{\theta} - df\wedge.\label{confn2}
\end{equation}
It follows
\begin{equation}
d_{k\theta'} = d_{k\theta} - k\cdot df\wedge.\label{confn2}
\end{equation}
Hence the map $I_f: \alpha \mapsto e^{kf} \alpha$ is an isomorphism between complexes $(F^* K^*, d_{k\theta})$ and
$(F^*K^*, d_{k\theta'})$, since
\begin{equation}
d_{k\theta'} ( e ^{kf }\alpha) = d_{k\theta} ( e^{kf} \alpha) + ( - k\cdot df) \wedge e ^{k\cdot f} \alpha = e^{k\cdot f }(d_{k\theta} \alpha).\label{conf1}
\end{equation}
It follows that the resulting terms $E^{p,q}_{k,0}$ are also conformal equivalent. Moreover, the map $I_f$ induces an isomorphism $I_f ^0$ between complexes
$$I_f ^0: (E^{p, q}_{k, 0}, d_{k, 0}) \to (E^{p, q}_{k, 0}, d_{k, 0} '), \, [\alpha] \mapsto [e^{kf}\alpha].$$
Inductively, this proves the first assertion of Theorem \ref{conf}.
Now we assume that ${\omega}' = {\omega} +d_\theta\rho$. Then $d_\theta {\omega}' = 0$. Using the injectivity of $L : {\Omega} ^1 (M ^{2n}) \to {\Omega} ^3 (M^{2n})$ we conclude that the Lee form of ${\omega}'$ is equal to the Lee form $\theta$ of ${\omega}$.
Denote by $L'$ the wedge product with ${\omega}'$, and by $[L']$ the induced operator on $H^*_l (M^{2n})$. Using ${\omega}' - {\omega} = d_\theta \rho$ and applying (\ref{comd2}), which implies that the wedge product with $d_\theta \rho$ maps $H^k_l (M^{2n})$ to zero, we conclude that the operators $L$ and $L'$ induce the same map $H^k_l (M^{2n}) \to H^{k+2}_{l+1} (M^{2n})$.
To prove the second assertion of Theorem \ref{conf} we use the following version of Five Lemma, whose proof is obvious and hence omitted.
\begin{lemma}\label{five}
Assume that the following diagram of vector spaces $A_i, B_i$ over a field ${\mathbb F}$ is commutative. If the rows are exact and $A_1 \to B_1$, $A_2 \to B_2$, $A_4 \to B_4$, $A_5 \to B_5$ are isomorphisms, then there is an isomorphism from $A_3$ to $B_3$, which also commutes with the other arrows.
$$
\xymatrix{A_1 \ar[r] \ar[d] & A_2 \ar[r]\ar[d] & A_3 \ar[r] & A_4 \ar[r]\ar[d] & A_5\ar[d]\\
B_1 \ar[r] & B_2 \ar[r] & B_3 \ar[r] & A_4 \ar[r] & A_B }
$$
\end{lemma}
The second assertion of Theorem \ref{conf} for $E^{p,q}_{k,1}$ follows immediately from the long exact sequence
(\ref{ex11}) and the formula (\ref{conf1}), if $0\le p \le q \le n-1$, taking into account Lemma \ref{fk1}.
To examine the term $E^{p, n}_{k,1}$, $0\le p \le n-1$, we need the following
\begin{lemma}\label{five2}
Assume that $\omega'=\omega+d_\theta\rho$. For $0\le p \le n$ there are linear maps $B^{n-p}_l:C^{n-p}_l(\omega)\rightarrow
C^{n-p}_l(\omega')$ such that the following two diagrams are commutative. (The symbol $I$ denotes the identity
mapping. The other mapping are defined in the proof.)
\xymatrix{
H^{n-(p+2)}_{l-1}(M^{2n})\ar[d]^I\ar[r]^{[L]} & C^{n-p}_l(\omega)\ar[d]^{B^{n-p}_l} &
C^{n-(p+1)}_{l-1}(\omega)\ar[d]^{B^{n-p-1}_{l-1}}\ar[r]^{[L^{p+1}]} & H^{n+p+1}_{l+p}(M^{2n})\ar[d]^I\\
H^{n-(p+2)}_{l-1}(M^{2n})\ar[r]^{[L']} & C^{n-p}_l(\omega') &
C^{n-(p+1)}_{l-1}(\omega')\ar[r]^{[L^{\prime p+1}]} & H^{n+p+1}_{l+p}(M^{2n})
}
\end{lemma}
\begin{proof}
Let us define first a linear mapping
$$
\tilde{B}^{n-p}_l:\ker d^-_l(\omega)\cap\Omega^{n-p}(M^{2n})\rightarrow\Omega^{n-p}(M^{2n}).
$$
Let $\eta\in\ker d^-_l\cap\Omega^{n-p}(M^{2n})$. This means that $d_l\eta=d_l^+\eta$ or equivalently
that $d_l\eta$ is primitive. The last assertion is again equivalent to the equality $\omega^p\wedge d_l\eta=0$. Since $(L')^p : {\Omega} ^{n-p}(M^{2n}) \to {\Omega} ^{n+p} (M^{2n})$ is an isomorphism, there is a unique $\eta'\in\Omega^{n-p}(M^{2n})$ such that
$$
\sum_{i=1}^p\binom{p}{i}\rho\wedge(d_\theta\rho)^{i-1}\wedge\omega^{p-i}\wedge d_l\eta=\omega^{\prime p}\wedge\eta'.
$$
Now we define $\tilde B^{n-p}_l$ by
$$\tilde{B}^{n-p}_l\eta:=\eta-\eta'.$$
We shall now prove that the element $d_l(\eta-\eta')$ is $\omega'$-primitive.
\begin{gather}
\omega^{\prime p}\wedge d_l(\eta-\eta')=-\omega^{\prime p}\wedge d_l\eta'+(\omega+d_1\rho)^p\wedge d_l\eta=\notag\\
=-d_{p+l}(\omega^{\prime p}\wedge\eta')+\omega^p\wedge d_l\eta+
\sum_{i=1}^p\binom{p}{i}(d_1\rho)^i\wedge\omega^{p-i}\wedge d_l\eta=\notag\\
=-d_{p+l}(\omega^{\prime
p}\wedge\eta')+d_{p+l}\bigg[\sum_{i=1}^p\binom{p}{i}\rho\wedge(d_1\rho)^{i-1}\wedge\omega^{p-i}
\wedge d_l\eta\bigg]=\notag\\
=d_{p+l}\bigg[-\omega^{\prime p}\wedge\eta'+\sum_{i=1}^p\binom{p}{i}\rho\wedge(d_{1}\rho)^{i-1}\wedge\omega^{p-i}
\wedge d_l\eta\bigg]=0.\notag
\end{gather}
In other words we have proved that $\tilde{B}^{n-p}_l$ maps $\ker d^-_l\cap\Omega^{n-p}(M^{2n})$ into $\ker (d')^-_l\cap\Omega^{n-p}(M^{2n})$, where $(d')^-_l$ is defined via the Lefschetz decomposition corresponding to ${\omega}'$.
Let us take now an element $\eta\in d_l\Omega^{n-p-1}(M^{2n})$, i. e. $\eta=d_l\gamma$, where $\gamma\in \Omega^{n-p-1}(M^{2n})$. Then
$d_l\eta=0$ and we have $\omega^{\prime p}\wedge\eta'=0$, which implies $\eta'=0$. We thus get $\tilde{B}^{n-p}_l\eta=\eta$.
Consequently we have proved that $\tilde{B}^{n-p}_l$ maps $d_l(\Omega^{n-p-1}(M^{2n}))$ into itself. Now it is obvious that $\tilde{B}^{n-p}_l$
induces a linear mapping
$$B^{n-p}_l:C^{n-p}_l(\omega)\rightarrow C^{n-p}_l(\omega').$$
Next we shall investigate the first diagram. First we define the mapping $[L]$. If $[\beta]\in H^{n-(p+2)}_{l-1}
(M^{2n})$, then we have an element $\beta\in\Omega^{n-(p+2)}(M^{2n})$ such that $d_{l-1}\beta=0$. Let us set
$$[L][\beta]:=[\omega\wedge\beta].$$
It is easy to see that this definition depends only on the cohomology class $[\beta]\in H^{n-p-2}_{l-1} (M^{2n})$. Namely, if $\tilde{\beta}=\beta+d_{l-1}\gamma$, then
$$
\omega\wedge(\beta+d_{l-1}\gamma)=\omega\wedge\beta+d_l(\omega\wedge\gamma),
$$
which shows that $[L][\tilde{\beta}]=[L][\beta]$. Similarly we define $[L']$. Let us take $[\beta]\in H^{n-(p+2)}_{l-1}
(M^{2n})$. Then $d_{l-1}\beta=0$, and we have
\begin{gather}
\sum_{i=1}^p\binom{p}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p-i}\wedge d_l(\omega\wedge\beta)=\notag\\
=\sum_{i=1}^p\binom{p}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p-i}\wedge(d_1\omega\wedge\beta+\omega\wedge
d_{l-1}\beta)=0.\notag
\end{gather}
We have $0=\omega^{\prime p}\wedge\eta'$, which implies $\eta'=0$. Thus we get $B^{n-p}_l[L][\beta]=
B^{n-p}_l[\omega\wedge\beta]=[\omega\wedge\beta-0]=[\omega\wedge\beta]$.
On the other hand we compute
\begin{gather}
[L'][\beta]=[\omega'\wedge\beta]=[(\omega+d_1\rho)\wedge\beta)]=[\omega\wedge\beta]+[d_1\rho\wedge\beta+\rho\wedge
d_{l-1}\beta]=\notag\\
=[\omega\wedge\beta]+[d_l(\rho\wedge\beta)]=[\omega\wedge\beta].\notag
\end{gather}
We have thus shown that that the first diagram is commutative.
We continue now with the second diagram. Again we define first the mapping $[L^{p+1}]$. For $[\eta]\in
C^{n-(p+1)}_{l-1}(\omega)$ there is a representative $\eta\in\Omega^{n-(p+1)}_{l-1}(M^{2n})$ such that $d^-_{l-1}\eta=0$.
This means that $d_{l-1}=d^+_{l-1}\eta$. We compute
\begin{gather}
d_{l+p}(\omega^{p+1}\wedge\eta)=d_{(p+1)+(l-1)}(\omega^{p+1}\wedge\eta)=d_{p+1}(\omega^{p+1})\wedge\eta+\omega^{p+1}
\wedge d_{l-1}\eta=\notag\\
=\omega^{p+1}\wedge d^+_{l-1}\eta=0.\notag
\end{gather}
The last term vanishes because the form $d^+_{l-1}\eta$ is primitive. Finally let us suppose that
$\eta=d_{l-1}\gamma$. Then we have
$$
\omega^{p+1}\wedge d_{l-1}\gamma=d_{p+1}(\omega^{p+1})\wedge\gamma+\omega^{p+1}\wedge d_{l-1}\gamma=d_{(p+1)+(l-1)}
(\omega^{p+1}\wedge\gamma).
$$
This shows that we can define $[L^{p+1}]$ by the formula
$$[L^{p+1}][\eta]:=[\omega^{p+1}\wedge\eta].$$
Now we are going
to prove the commutativity of the second diagram. For $[\eta]\in C^{n-(p+1)}_{l-1}(\omega)$ we have $B^{n-p-1}_{l-1}[\eta]=
[\eta-\eta']$, where $\eta'$ is uniquely determined by the equality
$$
\omega^{\prime p+1}\wedge\eta'=\sum_{i=1}^{p+1}\binom{p+1}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p+1-i}\wedge
d_{l-1}\eta.
$$
Further we have $[L']B^{n-p-1}_{l-1}[\eta]=[\omega'\wedge(\eta-\eta')]$. Now let us compute
\begin{gather}
\omega^{\prime p+1}\wedge(\eta-\eta')-\omega^{p+1}\wedge\eta=\omega^{\prime p+1}\wedge\eta
-\omega^{\prime p+1}\wedge\eta'-\omega^{p+1}\wedge\eta=\notag\\
=(\omega+d_1 \rho))^{p+1}\wedge\eta-\sum_{i=1}^{p+1}\binom{p+1}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p+1-i}\wedge
d_{l-1}\eta-\omega^{p+1}\wedge\eta=\notag\\
=\sum_{i=1}^{p+1}\binom{p+1}{i}(d_1 \rho)^i\wedge\omega^{p+1-i}\wedge\eta
-\sum_{i=1}^{p+1}\binom{p+1}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p+1-i}\wedge d_{l-1}\eta=\notag\\
=d_{p+1}\bigg(\sum_{i=1}^{p+1}\binom{p+1}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p+1-i}\bigg)\wedge\eta\notag\\
-\bigg(\sum_{i=1}^{p+1}\binom{p+1}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p+1-i}\bigg)\wedge d_{l-1}\eta=\notag\\
=d_{(p+1)+(l-1)}\bigg(\sum_{i=1}^{p+1}\binom{p+1}{i}\rho\wedge(d_1 \rho)^{i-1}\wedge\omega^{p+1-i}\wedge\eta\bigg).
\end{gather}
We have thus proved that $[L^{\prime p+1}]B^{n-p-1}_{l-1}[\eta]=[L^{p+1}][\eta]$.
\end{proof}
Clearly the second assertion of Proposition \ref{conf} for $E^{p,n}_{k,1}$, $0\le p \le n-1$, follows from Proposition \ref{prol1}, Lemma \ref{five}, and Lemma \ref{five2}.
Combining with (\ref{e13a}), which implies that $E^{n,n}_{k,1} = C^\infty (M^{2n})$, we obtain the second assertion of Proposition \ref{conf}.
This completes the proof of Theorem \ref{conf}.
\end{proof}
\begin{remark}\label{pois}
In \cite[Example 7.1]{FIL1998} the authors construct an example of a compact 6-dimensional nilmanifold $M^6$ equipped with a family of symplectic forms ${\omega}_t$, $t \in [0,1],$ with varying cohomology classes $[{\omega}_t] \in H^2 (M^6, {\mathbb R})$. They showed that the coeffective cohomology groups associated to ${\omega}_1$ and ${\omega}_2$ have different Betti number $b_4$. It follows that, using \cite[Lemma 2.7, Proposition 3.5, part II]{TY2009}, see also Remark \ref{his1}.1 above,
the $E_1$-terms of the associated spectral sequences for ${\omega}_0$ and ${\omega}_1$ are different.
\end{remark}
\section{The stabilization of the spectral sequences}
In this section we prove that the spectral sequences $\{E^{p,q}_{l,r} \}$ on l.c.s.manifolds
$(M^{2n}, {\omega}, \theta)$ converge to the Lichnerowicz-Novikov cohomology $H^*_l(M^{2n})$ at the second term $E^{*,*}_{l,2}$,
or at the $t$-term $E^{*,*}_{l,t}$ under some cohomological conditions posed on ${\omega}$ (Theorems \ref{stab1}, \ref{nil}, \ref{stcom}). As a consequence, we obtain a relation between the primitive cohomology groups and the de Rham cohomology groups of $(M^{2n}, {\omega})$, if $(M^{2n}, {\omega})$ is a compact symplectic manifold. This gives an answer to a question posed by Tseng and Yau in \cite{TY2009}, see Remark \ref{tsey}.
First we prove the following simple property of the second terms $E^{*,*}_{l,2}$, which will be used later in the proof of Theorem \ref{stcom}.
\begin{proposition}\label{symm} (cf. \cite[Proposizione 19]{Pietro2006}) Assume that $1\le p\le q \le n-1$. Then $E^{p, q}_{l, 2} = E^{p-1, q-1}_{l, 2}$.
\end{proposition}
\begin{proof} Let $\alpha \in E^{p, q}_{l+p, 1} =H^{q-p} ({\mathcal P}^* (M^{2n}), d^+_l) $ and $\tilde \alpha \in {\mathcal P}^{q-p} (M^{2n})$ its representative as in (\ref{e13}). The differential $d_{l+p,1}: E^{p,q}_{l+p, 1} \to E^{p+1, q}_{l+p, 1}$ is defined by
\begin{equation}
d_{l+p,1} (\alpha): = [d_{l+p} L^p \tilde \alpha] = [L^p d_{l} \tilde \alpha] \in E^{p+1, q}_{l+p, 1}.\label{stab1f}
\end{equation}
Using $d _{l+p} = d_{l+p}^+ + Ld_{l+p}^-$ and taking into account $d^+_{l} \tilde \alpha = 0$, we observe that
$[L^p d_l\tilde \alpha] \in E^{p+1, q}_{l+p, 1}$ has a representative $d_{l}^- (\tilde \alpha) \in {\mathcal P}^{q-p-1} (M^{2n})$ in $ H^{q-p-1} ({\mathcal P}^*( M^{2n}), d^+_{l-1}) =E^{p+1, q}_{l+p, 1} $, if $0\le p \le q \le n$. Equivalently, using (\ref{e13}), we rewrite $d_{l+p,1}$ for $0\le p \le q \le n$ as follows
\begin{equation}
d_{l+p,1} : H^{q-p} ({\mathcal P} ^*(M^{2n}), d^+_l) \to H^{q-p-1} ({\mathcal P}^*(M^{2n}), d^+_{l-1}), \, [\tilde \alpha] \mapsto [d_l^- \tilde \alpha]. \label{stab1a}
\end{equation}
Clearly Proposition \ref{symm} follows from (\ref{e13}) and the formula (\ref{stab1f}), (\ref{stab1a}).
\end{proof}
Now assume that ${\omega} = d_k\tau$ for some $k \in \Z$ and $\tau \in {\Omega}^1(M^{2n})$. Since $d_1 {\omega} = d_k {\omega} = 0$, it follows that $(k-1)\theta \wedge {\omega} =0$. Since $L$ is injective, we get $k = 1$. The following theorem is a generalization of \cite[Theorem 2]{PV2006} for the symplectic case $\theta = 0$.
\begin{theorem}\label{stab1} (cf. \cite[Theorem 2]{PV2006}) Assume that ${\omega} = d_1 \tau$. Then $E^{p,q}_{l,2} = 0$, if $1\le p \le q \le n-1$.
If $0\le q \le n$, then $E^{0,q}_{l,2} = H^q_l (M^{2n})$. If $ 0\le p \le n$ then $E^{p,n}_{l+p,2} = H^{n+p}_{l+p} (M^{2n})$. Thus the spectral sequence $\{ E^{p,q}_{l,r}, d_{l,r}\}$ stabilizes at the term $E_{l,2}$.
\end{theorem}
\begin{proof} Assume that ${\omega}= d_1 \tau$. Then for any $d_{l-1}$-closed form $\alpha$ we have
\begin{equation}
d_1\tau \wedge \alpha = d_{l}(\tau \wedge \alpha). \label{exact0}
\end{equation}
Hence the induced operator $\bar L$ in the exact sequence (\ref{ex11}) satisfies
\begin{equation}
\bar L( H_{l-1} ^{q- (p+1)} (M^{2n})) = 0 \in H_l ^{q+1 -p} (M^{2n}).
\label{exact0a}
\end{equation}
The equality (\ref{exact0a}) and the exact sequence (\ref{ex11}) lead to the following exact sequence for $0\le p \le q \le n-1$.
\begin{equation}
0 \to H^{q-p} _l(M^{2n}) \stackrel{\bar L^p}{\to} E^{p,q}_{l+p,1} \stackrel{\delta_{p,q}}{\to} H^{q-(p+1)} _{l-1}(M^{2n}) \to 0.\label{exact1}
\end{equation}
Using the isomorphism $E^{p,q}_{l+p, 1} = H^{q-p}({\mathcal P}^* (M^{2n}), d_l ^+)$ and the formulae (\ref{delta1}) and (\ref{lpex}) describing $\delta_{p,q}$ and $\bar L^p$, we rewrite the
exact sequence (\ref{exact1}) as follows
\begin{equation}
0 \to H^{q-p} _l(M^{2n}) \stackrel{[\Pi_{pr}]}{\to} H^{q-p}({\mathcal P}^*(M^{2n}), d_l ^+) \stackrel{[d^-_l]}{\to} H^{q-(p+1)} _{l-1}(M^{2n}) \to 0.\label{exact1a}
\end{equation}
The proof of the first and second assertion of Theorem \ref{stab1} is based on our analysis of the long exact sequence of cohomology groups
associated with the short exact sequence (\ref{exact1a}). By (\ref{stab1a}), for $ 0 \le p \le q \le n-1$, the differential $d_{l+p,1} : E^{p, q}_{l+p, 1} \to E^{p+1, q}_{l+p,1}$ induces the following boundary
operator
\begin{equation}
\hat d ^-_l : H^{q-p}({\mathcal P}^*(M^{2n}), d_l^+) \to H ^{q-p-1} ({\mathcal P}^*(M^{2n}), d_{l-1} ^+), \, [\tilde \alpha]\mapsto [d^-_l \tilde \alpha], \label{exact1a1}
\end{equation}
for $\tilde \alpha \in {\mathcal P} ^{q-p} (M^{2n})$.
\begin{lemma}\label{id1a} The short exact sequence (\ref{exact1a}) generates a short exact sequence of the following chain complexes for $1\le p \le q \le n-1$:
\begin{eqnarray}
0 \to (H^{q-p}_l (M^{2n}), \tilde d_l : = 0) \stackrel{[\Pi_{pr}]}{\to} (H^{q-p}({\mathcal P}^*(M^{2n}), d_l^+), \hat d_l^-)\to \nonumber\\
\stackrel{[d^-_l]}{\to} (H^{q-(p+1)} _{l-1} (M^{2n}), \tilde d_{l-1} : = 0) \to 0.\label{id1a1}
\end{eqnarray}
\end{lemma}
\begin{proof} It is useful to rewrite the sequence (\ref{id1a1}) of chain complexes as the following diagram
$$
\xymatrix{
H^{q-(p-1)}_{l+1}(M^{2n})\ar[r] ^ {[\Pi_{pr}]} \ar[d] ^{\tilde d_{l+1} =0} & H^{q-(p-1)}({\mathcal P}^*(M^{2n}), d_{l+1} ^+) \ar[r] ^{[d_{l+1}^-]} \ar[d]^{\hat d_{l+1} ^-} & H^{q-p}_{l}(M^{2n})\ar[r] \ar[d] ^{\tilde d_l =0} &\\
H^{q-p}_{l}(M^{2n})\ar[r] ^ {[\Pi_{pr}]} \ar[d] ^{\tilde d_l =0} & H^{q-p}({\mathcal P}^*(M^{2n}), d_{l} ^+) \ar[r] ^{[d_{l}^-]} \ar[d]^{\hat d_l ^-} & H^{q-(p+1)}_{l-1} (M^{2n})\ar[r] \ar[d] ^{\tilde d_{l-1} =0 } &\\
H^{q-(p+1)}_{l-1}(M^{2n})\ar[r] ^ {[\Pi_{pr}]} & H^{q-(p+1)}({\mathcal P}^*(M^{2n}), d_{l-1} ^+) \ar[r] ^{[d_{l-1}^-]} & H^{q-(p+2)}_{l-2}(M^{2n}) \ar[r] &
}
$$
To prove Lemma \ref{id1a}, it suffices to show that the above diagram is commutative, or equivalently
\begin{eqnarray}
\, \hat d^-_l[\Pi_{pr}] = \tilde d_l [\Pi_{pr}] = 0, \label{newcoh1a}\\
\, [d^-_{l-1}]\hat d^-_l = \tilde d_{l-1} [d^-_l] = 0.\label{newcoh2a}
\end{eqnarray}
Let $\alpha \in H^{q-p}_l(M^{2n})$ and $\tilde \alpha \in {\Omega}^{q-p}(M^{2n})$ its representative. Let $\tilde \alpha = \tilde \alpha _{pr} + L \tilde \beta _{pr} + L^2 \gamma$ be the Lefschetz decomposition of $\tilde \alpha$. Using $d_l\tilde \alpha = d^+_l\tilde \alpha = d^+_{l} \tilde \alpha _{pr} =0$ we obtain $L d^- _l \tilde \alpha _{pr} = L d^+_{l-1}\tilde \beta _{pr} + L^ 2 (d^-_{l-1} \tilde \beta _{pr} + d_{l-2} \gamma)$. Hence
$$
0=L(d_l^-\tilde{\alpha}_{pr}+d_{l-1}^+\tilde{\beta}_{pr})+L^2(d_{l-1}^-\tilde{\beta}_{pr}+
d_{l-2}\gamma),
$$
which implies $d_l^-\tilde{\alpha}_{pr}+d_{l-1}^+\tilde{\beta}_{pr}=0$ thanks to the uniqueness of the second Lefschetz decomposition. It follows
\begin{equation}
\hat{d}_l^-[\Pi_{pr}]\alpha=\hat{d}_l^-[\tilde{\alpha}_{pr}]=[d_l^-\tilde{\alpha}_{pr}]=-[d_{l-1}^+\tilde{\beta}_{pr}]
=0\in H^{q-(p+1)}({\mathcal P}^*(M^{2n}), d^+_{l-1}). \label{stab63a}
\end{equation}
Let $\beta \in H^{q-p}({\mathcal P}^*(M^{2n}), d_l^+) $ and $\tilde \beta \in {\mathcal P} ^{q-p}(M^{2n})$ its representative. Then
\begin{equation}
\, [ d_{l-1}^- ] \hat d^-_{l} \beta = [d_{l-1} ^- d^-_{l} \tilde \beta] = 0 \in H^{q-(p+2)}_{l-2}(M^{2n}). \label{stab64a}
\end{equation}
Clearly (\ref{newcoh1}) and (\ref{newcoh2}) follow from (\ref{stab63}) and (\ref{stab64a}). This completes the proof of Lemma \ref{id1a}.
\end{proof}
The short exact sequence (\ref{id1a1}) in Lemma \ref{id1a} generates the following associated long exact sequence of the cohomology groups
\begin{equation}
\to E^{p-1, q}_{l+p, 2} \to H^{q-p}_{l}(M^{2n}) \stackrel{\delta} \to H^{q-p}_{l}(M^{2n}) \to E^{p, q}_{l+p, 2} \to H^{q-(p+1)}_{l-1}(M^{2n}) \stackrel{\delta} {\to} \label{stab11}
\end{equation}
where $\delta$ is the connecting homomorphism.
\begin{lemma}\label{deltata} We have $\delta (x) = x$ for all $x \in H^{q-(p+1)}_{l-1}(M^{2n})$ and for all $1\le p \le n-1$.
\end{lemma}
\begin{proof}[Proof of Lemma \ref{deltat}] Let $x \in H^{q-(p+1)}_{l-1}(M^{2n})$. Using (\ref{id1a1}) we write
$x= [d^-_l] y, y \in H^{q-p}({\mathcal P} ^*( M^{2n}),d^+_l )$. By definition of the connecting homomorphism
we have $\delta x = [\hat d^-_l y] = x$. This completes the proof of Lemma \ref{deltata}.
\end{proof}
Clearly Lemma \ref{deltata} implies the first assertion of Theorem \ref{stab1}.
Now let us consider the case $ p = 0,\, q\le n-1$. Then
$ E^{-1, q}_{l,1} = 0$.
The previous short exact sequence (\ref{id1a1}) of chain complexes is now replaced by the new
one, where the cohomology groups on the line containing $E^{-1, q}_{l,1}$ left and right to $E^{-1, q}_{l,1}$ are zero. Let us write the new short exact sequence explicitly
as the following commutative diagram
$$
\xymatrix{
0\ar[r] ^ {0} \ar[d] ^{0} & 0 \ar[r] ^{0} \ar[d]^{0} & 0 \ar[r] \ar[d] ^{0} &\\
H^{q}_{l}(M^{2n})\ar[r] ^ {[\Pi_{pr}]} \ar[d] ^{\tilde d_l =0} & H^{q}({\mathcal P}^*(M^{2n}), d_{l} ^+) \ar[r] ^{[d_{l}^-]} \ar[d]^{\hat d_l ^-} & H^{q-1}_{l-1} (M^{2n})\ar[r] \ar[d] ^{\tilde d_{l-1} =0 } &\\
H^{q-1}_{l-1}(M^{2n})\ar[r] ^ {[\Pi_{pr}]} & H^{q-1}({\mathcal P}^*(M^{2n}), d_{l-1} ^+) \ar[r] ^{[d_{l-1}^-]} & H^{q-2}_{l-2}(M^{2n}) \ar[r] &
}
$$
The resulting exact sequence of the cohomology groups now are
$$ 0 \to H^q _{l}(M^{2n}) \to E^{0, q}_{l, 2} \to H ^{q-1}_{l-1}(M^{2n}) \stackrel{ \delta} {\to } H ^ {q- 1} _{l -1}(M^{2n})...$$
Since $ \delta = Id$, we obtain
\begin{equation}
E^{0, q}_{l,2} = H^{q}_l (M^{2n}), \label{1411b}
\end{equation}
which proves the second assertion of Theorem \ref{stab1}.
Next we compute $E^{p, n}_{l, 2} = E^{p, n}_{l, 1} / d_{l,1} (E^{p-1, n}_{l,1})$ for $0\le p \le n-1$. Since ${\omega} = d_1 \tau$, the map
$[L] : H^{n -(p +2)}_{l-1} (M^{2n}) \to C^{n-p}_l$ sends $[\alpha]$ to $[d_l (\tau \wedge \tilde\alpha)] = 0 \in C^{n-p}_l$. Thus Proposition \ref{prol1} implies that the following sequence is exact for $0\le p \le n-1$
\begin{equation}
0 \to C^{n-p}_l \stackrel{[L^p]}{\to}E^{p,n}_{l+p, 1} \stackrel{\delta_{p,n}}{\to} C^{n-(p+1)}_{l-1} \stackrel{[ L^{p+1}]}{\to} H^{n+p+1}_{l+p} (M^{2n})\to 0.\label{stab5}
\end{equation}
Set for $ -1\le p \le n-1$
\begin{equation}
T^{n-(p+1)}_{l-1} : = \ker [L^{p+1}]: C^{n-(p+1)} _{l-1} \to H^{n+p+1}_{l+p} (M^{2n}).\label{deft}
\end{equation}
Then we obtain from the exact sequence (\ref{stab5}) the following short exact sequence
\begin{equation}
0 \to C^{n-p}_l \stackrel{[L^p]}{\to}E^{p,n}_{l+p, 1} \stackrel{\delta_{p,n}}{\to} T^{n-(p+1)}_{l-1}\to 0.\label{stab6}
\end{equation}
Using the isomorphism $E^{p,n}_{l+p, 1} = {\mathcal P} ^{n-p} (M^{2n}) / d^+_l ({\mathcal P}^{n-p-1}(M^{2n}))$ and the formulas (\ref{new1a}) and (\ref{deltac}) describing $[L^p]$ and $\delta_{p,n}$
in the exact sequence (\ref{prol1a}) of Proposition \ref{prol1}, we rewrite the short exact sequence (\ref{stab6}) as follows
\begin{equation}
0 \to C^{n-p}_l \stackrel{ [\Pi_{pr}]}{\to} \frac{{\mathcal P} ^{n-p}(M^{2n})}{ d^+_l ({\mathcal P}^{n-p-1}(M^{2n}))} \stackrel{[d_l ^-]}{\to} T^{n-(p+1)}_{l-1} \to 0. \label{stab61}
\end{equation}
Recall that the map $[\Pi_{pr}]$ is already defined in section 4. It is the quotient map of the map $\Pi_{pr}: (\ker d_l ^- \cap {\Omega} ^{n-p} (M^{2n})) \to {\mathcal P} ^{n-p} (M^{2n})$, see the explanation of (\ref{new1a}).
Next recall that the map $[d_l^- ]$ is the quotient map of the map $d_l ^-: {\mathcal P} ^{n-p}(M^{2n}) \to {\mathcal P} ^{n-p-1}(M^{2n}) \cap \ker d^-_{l-1}$, see the explanation of (\ref{deltac}). (We now explain why this map is also well-defined in (\ref{stab61}). First we have $d^-_l (d ^+ _l \gamma) = d^+_{l-1} d^-_l \gamma = d_{l-1} (d^-_l \gamma) = 0 \in C^{n-(p+1)}_{l-1}$. Furthermore for $\alpha \in {\mathcal P} ^{n -p} (M^{2n})$
$$ L^{p+1} d_l^- (\alpha) = L ^p ( Ld_l ^- \alpha ) = L^pd_l \alpha = d_{l +p} L^p \alpha = 0 \in H^{n +p +1} _{ l+p } (M ^{2n}).$$
Hence
$$[d_l^-] (\frac{{\mathcal P} ^{n-p}(M^{2n})}{ d^+_l ({\mathcal P}^{n-p-1}(M^{2n}))})\subset T ^{n -(p +1)}_{l-1} = \ker [L^{p+1}].$$
Thus $[d_l^-]$ is well-defined.)
Note that the differential $d_{l+p, 1} : E^{p,n}_{l+p, 1} \to E^{p+1, n}_{l+p, 1}$ induces the following boundary operator
\begin{equation}
\hat d^-_l: \frac{{\mathcal P} ^{n-p}(M^{2n})}{ d^+ _l({\mathcal P}^{n-p-1}(M^{2n}))} \to \frac{{\mathcal P} ^{n-p-1}( M^{2n})}{ d^+_{l-1} ({\mathcal P}^{n-p-2}(M^{2n}))}, [\tilde \alpha] \mapsto [d^-_l \tilde \alpha], \label{stab62}
\end{equation}
for $\tilde \alpha \in {\mathcal P}^{n-p}(M^{2n})$. The map $\hat d^-_l$ is well-defined, since by (\ref{com1a}) $d^-_l d^+_l \alpha = d^+_{l-1} d ^- _l \alpha$ for $\alpha \in {\mathcal P}^{n-p-1}(M^{2n})$.
\begin{lemma}\label{newcoh} The short exact sequence (\ref{stab61}) generates a short exact sequence of the following chain complexes
\begin{equation}
\xymatrix{
0 \ar[r] & C^{n-p}_l \ar[d]^{\tilde d_{l}:= 0} \ar[r]^{ [\Pi_{pr}]} & \frac{{\mathcal P} ^{n-p} (M^{2n})}{ d^+_l ({\mathcal P}^{n-(p+1)}(M^{2n}))}\ar[d]^{ \hat d^-_l} \ar[r]^{[d_l ^-]}& T^{n-(p+1)}_{l-1}\ar[d]^{ \bar d_{l-1}: = 0}\ar[r] & 0\\
0 \ar[r] & C^{n-(p+1)}_l \ar[r]^{ [\Pi_{pr}]} & \frac{{\mathcal P} ^{n-(p+1)} (M^{2n})}{ d^+_l ({\mathcal P}^{n-(p+2)}(M^{2n}))}\ar[r]^{[d_l ^-]} & T^{n-(p+2)}_{l-1}\ar[r]& 0
}
\label{stab6a}
\end{equation}
\end{lemma}
\begin{proof}[Proof of Lemma \ref{newcoh}] It suffices to show that
\begin{eqnarray}
\, \hat d^-_l[\Pi_{pr}] = \tilde d_l \Pi_{pr} = 0, \label{newcoh1}\\
\, [d^-_{l-1}]\hat d^-_l = \bar d_{l-1} [d^-_l] = 0.\label{newcoh2}
\end{eqnarray}
Let $\alpha \in C^{n-p}_l$ and $\tilde \alpha \in {\Omega}^{n-p}(M^{2n})$ its representative. Let $\tilde \alpha = \tilde \alpha _{pr} + L \tilde \beta _{pr} + L^2 \gamma$ be the Lefschetz decomposition of $\tilde \alpha$. Using $d_l\tilde \alpha = d^+_l\tilde \alpha$ we obtain $L d^- _l \tilde \alpha _{pr} = L d^+_{l-1}\tilde \beta _{pr} + L^ 2 (d^-_{l-1} \tilde \beta _{pr} + d_{l-2} \gamma)$. Hence
\begin{equation}
\hat d^-_l [ \Pi_{pr}]\alpha= [d^-_l \tilde \alpha _{pr}]= [d_{l-1}^+ \tilde \beta _{pr}] = 0 \in \frac{{\mathcal P} ^{n-p-1}( M^{2n})}{ d^+_{l-1} ({\mathcal P}^{n-p-2}(M^{2n}))}.\label{stab63}
\end{equation}
Let $\beta \in ({\mathcal P} ^{n-p}(M^{2n}))/ d^+_l ({\mathcal P}^{n-p-1}(M^{2n}))$ and $\tilde \beta \in {\mathcal P} ^{n-p}(M^{2n})$ its representative. Then
\begin{equation}
\, [ d_{l-1}^- ] \hat d^-_{l} [\beta] = [d_{l-1} ^- d^-_{l} \tilde \beta] = 0. \label{stab64}
\end{equation}
Clearly (\ref{newcoh1}) and (\ref{newcoh2}) follow from (\ref{stab63}) and (\ref{stab64}). This completes the proof of Lemma \ref{newcoh}.
\end{proof}
The short exact sequence (\ref{stab6a}) in Lemma \ref{newcoh} generates the following associated long exact sequence of the cohomology groups
\begin{equation}
E^{p-1,n}_{l+p,2} \to T^{n-p}_{l} \stackrel{\delta} \to C^{n-p}_{l} \to E^{p, n}_{l+p, 2} \to T^{n-(p+1)}_{l-1} \stackrel{\delta} {\to} C^{n-(p+1)}_{l-1} \to , \label{stab11}
\end{equation}
where $\delta$ is the connecting homomorphism.
\begin{lemma}\label{deltat} We have $\delta (x) = x$ for all $x \in T^{n-(p+1)}_{l-1}$ and for all $0\le p \le n-1$.
\end{lemma}
\begin{proof}[Proof of Lemma \ref{deltat}] Let $x \in T^{n-(p+1)}_{l-1}$. Using (\ref{stab6a}) we write
$x= [d^-_l] y, y \in ({\mathcal P} ^{n-p}( M^{2n}))/ d^+_l ({\mathcal P}^{n-p-1}(M^{2n}))$. By definition of the connecting homomorphism
we have $\delta x = [\hat d^-_l y] = x$. This completes the proof of Lemma \ref{deltat}.
\end{proof}
Now let us complete the proof of Theorem \ref{stab1}. From Lemma \ref{deltat} and the long exact sequence (\ref{stab11}) we obtain $E^{p,n}_{l+p, 2} = C^{n-p}_{l}/ T ^{n-p}_{l}$. Taking into account (\ref{deft}) which defines $T^{n-p}_{l}$ to be the kernel of the surjective homomorphism
$[L^{p}]: C^{n-p}_{l} \to H^{n+p}_{l+p}(M^{2n})$, we conclude that
\begin{equation}
E^{p,n}_{l+p, 2} = H^{n+p}_{l+p}(M^{2n})\text { for } 0\le p \le n-1.\label{stab22n}
\end{equation}
Next, by Lemma \ref{fk1} $E^{n,n}_{l, 1} = C^\infty (M^{2n})$. Since $d_{l, 1} (E^{n,n}_{l,1}) = 0$, using (\ref{stab62}) we get
\begin{equation}
E^{n, n}_{l+n, 2} = \frac{ C^\infty (M^{2n})}{d_l ^- ({\mathcal P} ^1 (M^{2n}))} = H ^0 ({\mathcal P} ^* (M^{2n}), d_l ^-).\label{stab22p}
\end{equation}
By Proposition \ref{primi}, $ d_l ^- $ is proportional to $ (d_l)^*_{\omega}$. Applying the symplectic star operator we get from (\ref{stab22p})
\begin{equation}
E^{n, n}_{l+n, 2}= H ^0 ({\mathcal P} ^* (M^{2n}), d_l ^-)= H^0 ({\mathcal P}^*(M^{2n}), (d_l) ^* _{\omega}) = H^{2n}_{n +l} (M^{2n}).\label{stab22m}
\end{equation}
Clearly
the third assertion of Theorem \ref{stab1} follows from (\ref{stab22n}) and (\ref{stab22m}). The last assertion of Theorem \ref{stab1} follows immediately.
This completes the proof of Theorem \ref{stab1}.
\end{proof}
From the exact sequence (\ref{exact1}) we obtain immediately the following
\begin{corollary} Assume that ${\omega} = d_1 \tau$. For $0\le p \le q\le n-1$ we have
\begin{equation}
E^{p, q}_{l+p, 1} = H^{q-p}_l (M^{2n}) \oplus H^{q-p-1}_{l-1} (M^{2n}).\label{stab12}
\end{equation}
\end{corollary}
Theorem \ref{stab1} can be generalized as follows. Assume that ${\omega} ^p = d_T \rho$ for some $\rho \in {\Omega}^{2p-1} (M^{2n})$, in particular $d_T {\omega} ^p = 0$. Clearly $d_p ({\omega} ^p ) = 0 = d_T ({\omega} ^p)$ implies that $ T = p$, since $L ^p : {\Omega}^1 (M^{2n}) \to {\Omega} ^{1+2p} (M^{2n})$ is injective. Furthermore we have
${\omega}^{k+T} = d_{k +T}(\rho \wedge {\omega})$ for all $ t\ge 0$. Thus there exists a minimal number $T$ such that
${\omega} ^T = d_T \rho$ for some $\rho \in {\Omega}^{2T-1} (M^{2n})$.
\begin{theorem}\label{nil} (cf. \cite[Theorem 3]{PV2006}) Assume that ${\omega} ^T = d_T \rho$ and $T \ge 2$. Then the spectral sequence $(E^{p, q}_{l, r}, d_{l, r})$ stabilizes at the term $E^{*,*}_{l, T+1}$.
\end{theorem}
\begin{proof}
Our proof of Theorem \ref{nil} exploits the construction of the exact couple associated with a filtered complex. We use many ideas from \cite{Pietro2006}. The main idea is to find a short exact sequence, whose middle term is $E^{*,*}_{l, T}$, and moreover, this short exact sequence is induced from the trivial action of the operator $L ^T$ on (a part of) complexes entering in the derived exact couples (cf. with the proof of Theorem \ref{stab1}). The condition $T \ge 2$ is necessary for Lemma \ref{isom1} below.
Let us begin with recalling the construction of the derived exact couple associated with a
filtered complex $(F^pK^*, d_l)$ following \cite[p.37-43]{McCleary2001}.
We associate with a filtration $(F^pK^*, d_l)$ the
following exact couple
\begin{equation}\xymatrix{ D^{p+1, *}_l \ar[r]^i & D^{p, *}_l \ar[d]^j\\
& E^{p, *}_{l, 1}\ar[lu]^\delta },\label{excp}
\end{equation}
where $D^{p, q}_l: = H^{p+q} (F^pK^*, d_l)$, which we also abbreviate as $H^{p+q}_l (F^p K^*)$, and
$$\to D^{p+1, q-1}_l \stackrel{i}{\to }D^{p,q}_l\stackrel{j}{\to} E^{p, q}_{l,1}\stackrel{\delta} {\to} D^{p+1, q}_l\stackrel{i}{\to} D^{p, q +1}_{l,1} \stackrel{j}{\to}E^{p,q+1}_{l,1} \to$$
is the long exact sequence of cohomology groups associated with the following short exact sequence of chain complexes
\begin{equation}
0 \to (F^{p+1} K^{p+q}, d_l) \stackrel{\tilde i}{\to} (F^{p} K^{p+q}, d_l) \stackrel{\tilde j}{\to } (E^{p, q} _{l, 0}, d_{l,0}) \to 0.
\end{equation}
The differential $d_{l, 1}:E^{p, q}_{l, 1} \to E^{p+1, q}_{l, 1}$, defined in (\ref{stab1f}), satisfies the following relation: $d_{l, 1} = j\circ \delta$.
We refer the reader to \cite{McCleary2001} for a comprehensive exposition on the relation between the spectral sequence of a filtration and its associated exact couple.
Set $(D^{*,*} _{l})^0 : = D^{*,*} _{l}$. We define the $t$-th derived exact couple of the exact couple (\ref{excp}), $t\ge 1$,
\begin{equation}
(D_l^{p+1,q-1})^{(t)}\stackrel{i^{(t)}}{\to}(D_l^{p-t, q+t})^{(t)}\stackrel{j ^{(t)}}{\to} E^{p,q}_{l,t+1}\stackrel{\delta^{(t)}}{\to}(D_l^{p+1,q})^{(t)}\label{dct}
\end{equation}
inductively as follows \cite[p. 38]{McCleary2001}.
\begin{eqnarray}
(D_l^{p,q}) ^{(t)}: = i (D_l^{p+1, q-1})^{(t-1)} \subset D_{l} ^{p, q}, \label{excdt}\\
i ^{(t)} (i ^t x) : = i (i^t x) \text { for } i ^t x \in (D_l^{p, q}) ^{(t)}, \label {excit}\\
E^{p, q}_{l, t+1} : = \frac{\ker d_{l, t} \cap E^{p, q}_{l, t}}{d_{l, t-1} (E^{p -t+1, q+t -2}_{l, t})}, \label{exet}\\
j ^{(t)} (i ^t x): = [j^{(t-1)} \circ (i^{(t-1)} x )], \label{excjt}\\
\delta ^{(t)} ([e]): = \delta ^{(t-1)} (e) \in i (D_l^{p, q})^{(t-1)},\\
d_{l, t+1} : = j ^{(t)} \circ \delta ^{(t)}.
\end{eqnarray}
Next we consider the following commutative diagram
\begin{equation}
\xymatrix{D^{0,q-p}_l \ar[r]^{ \bar L}\ar[d]^{\bar L^p} & D^{0, q-p+2}_{l+1}\ar[d] ^{\bar L^{p-1}}\\
D^{p, q}_{l+p} \ar[r] ^{i} & D^{p-1, q+1}_{l+p}
} \label{excp1}
\end{equation}
where $\bar L ^p$ is induced by the linear operator $L^p: {\Omega} ^{q-p} (M^{2n}) \to {\Omega} ^{q+p} (M^{2n})$.
The diagram (\ref{excp1}) leads us to consider the following diagram
\begin{equation}
\xymatrix{(D^{0,q-p}_l) ^{(t)} \ar[r]^{\bar L^{(t)}}\ar[d]^{(\bar L^p) } & (D^{0, q-p+2}_{l+1})^{(t)}\ar[d] ^{(\bar L^{p-1}) }\\
(D^{p, q}_{l+p})^{(t)} \ar[r] ^{i^{(t)}} & (D^{p-1, q+1}_{l+p}) ^{(t)}
}\label{excpk}
\end{equation}
where $\bar L^{(t)}$ (resp. $(\bar L ^p)$) is the restriction of $\bar L$ (resp. $\bar L^p$) to $(D^{p, q}_{l+p})^{(t)}$.
\begin{lemma}\label{isom1} For $t \ge 1$ and $p\ge 1$, $T\ge 2$, the following statements hold.\\
1. The diagram (\ref{excpk}) is commutative.\\
2. $(\bar L^{p}) $ is an isomorphism.\\
3. ${\rm Im} \, (i ^{(t)}) = {\rm Im} \,( (\bar L ^{p-1}))$. \\
4. $(\bar L^{(T-1)}) (D^{0,q-p}_{l}) ^{(T-1)} = 0$, if $d_T{\omega} ^T = 0$.
\end{lemma}
\begin{proof} 1. The commutativity of (\ref{excpk}) is an immediate consequence of the commutativity of the diagram
(\ref{excp1}).
2. We prove the second assertion of Lemma \ref{isom1} by induction, beginning with $t = 1$. Let $x = i (\alpha) \in (D^{p, q} _{l+p}) ^{(1)} = i (D^{p+1, q-1}_{l+p}) = i (H^{p+q}_{l+p}(F^{p+1}K^*)) \subset D^{p,q}_{l+p}$. Then there is an element $\alpha' \in {\Omega} ^{q-p-2} (M^{2n})$ such that $[L^{p+1}\alpha'] = \alpha \in D^{p+1, q-1}_{l+p}$, or equivalently $d_{l+p} (L^{p+1} \alpha') = 0$. Hence $L^{p+1} (d_{l-1} (\alpha')) = 0$. Since $d_{l-1}\alpha ' \in {\Omega} ^{q-p-1}(M^{2n})$, $L^{p+1} (d_{l-1} (\alpha')) = 0$ implies that $d_{l-1} \alpha ' = 0$, so $\alpha' \in H^{q-p-2}_{l-1} (M^{2n})$.
Hence $x = \bar L^p ( L \alpha')$, and $L\alpha ' = i (\alpha') \in (D^{0, q-p}_l )^{(1)}$. This proves that the linear map $(\bar L^{p})$
is surjective for $t=1$. Furthermore, the map $(\bar L^{p})$ is injective for $t =1$, since $L^p: {\Omega}^{q-p}M^{2n} \to {\Omega} ^{q+p}M^{2n}$ is injective, and $L^p d_{l} = d_{l+p} L^p$. This proves Lemma \ref{isom1}.2 for $t =1$.
Now assume that Lemma \ref{isom1}.2 has been proved for $t = k$. Since $(D^{0, q-p}_l) ^{(k+1)}$ is a subset
of $(D^{0, q-p}_l) ^{(k)}$ the injectivity of $(\bar L^{p})$ follows from the inductive statement.
The surjectivity of $(\bar L^p)$ also follows from the commutativity of the diagram (\ref{excpk}), which implies
that $(\bar L^{p-1})$ maps the image $ i (( D^{0,q-p}_l)^{(k)})$ onto the set $i^{(k)} ( D^{p, q}_{l+p}) ^{(k)}
= (D^{p-1, q+1}_{l+p})^{(k+1)}$.
This proves Lemma \ref{isom1}.2 for all $t\ge 2$.
3. Clearly Lemma \ref{isom1}.3 is a consequence of Lemma \ref{isom1}.2 and the commutativity of the diagram (\ref{excpk}).
4. Let us compute $\bar L ^{(T-1)}\beta$ for $\beta \in (D^{0, q-p-2}_l)^{(T-1)}$, where $T\ge 2$. By definition
$\bar L ^{(T-1)} \beta = \bar L (i^{(T-1)}[\tilde \beta])$, where $\tilde \beta = L^{T-1}\hat \beta\in {\Omega}^{q-p-2}(M^{2n}) $ for some $\hat \beta \in {\Omega}^{q-p - 2T}(M^{2n})$, and $[\tilde \beta] \in D^{0,q-p}_l$, in particular we have $d_l \tilde \beta = 0$. Thus $L^{T-1}d_{l-T-1} \tilde \beta = 0$. Since $L^{T-1}: {\Omega}^{q-p-2T +2 } (M^{2n}) \to {\Omega}^{q-p}(M^{2n})$ is injective we get $d_{l-T +1} \hat \beta = 0$. Now we have
\begin{equation}
\bar L^{(T-1)} \beta = \bar L ( i ^{(T-1)}\tilde \beta) = i^{(T-1)}([ L^T \hat \beta])\in (D^{p-1, q+1})^{(T)}_{l+p}, \label{isom14}
\end{equation}
by Lemma \ref{isom1}.1.
Note that
\begin{equation}
[L^T \hat \beta] = [d_{T} \rho \wedge \hat \beta] = [d_{T +(l-T+1)} ( \rho \wedge \hat \beta)] = 0\in D^{0, q-p +2}_{l+1}. \label{isom14a}
\end{equation}
Clearly (\ref{isom14}) and (\ref{isom14a}) imply the last assertion of Lemma \ref{isom1}.
\end{proof}
Lemma \ref{isom1}.4 and the $(T-1)$th-derived exact couple yield the following short exact sequence
\begin{equation}
0 \to (D^{p-(T-1), q+(T-1)}_l) ^{(T-1)} \stackrel{j^{(T-1)}}{\to} E^{p, q}_{l,T} \stackrel{\delta^{(T-1)}}{\to} (D^{p+1, q}_l) ^{(T-1)} \to 0.\label{vani2}
\end{equation}
\begin{lemma}\label{complex} The short exact sequence (\ref{vani2}) generates a short exact sequence of the following chain complexes
$$
\xymatrix{
0 \ar[r] & (D^{p-(T-1), q+(T-1)}_l) ^{(T-1)} \ar[r]_{j^{(T-1)}}\ar[d]^{\tilde d_{l, T} = 0} & E^{p, q}_{l,T} \ar[r]_{\delta^{(T-1)}}\ar[d] ^{d_{l, T}} & (D^{p+1, q}_l) ^{(T-1)} \ar[d]^{\tilde d_{l,T} = 0}\ar[r] & 0 \\
0 \ar[r] & (D^{p+1, q}_l) ^{(T-1)} \ar[r]_{j^{(T-1)}} & E^{p+ T, q-T +1}_{l,T} \ar[r]_{\delta^{(T-1)}}& (D^{p+T+1, q-T +1}_l) ^{(T-1)} \ar[r] & 0
}
$$
\end{lemma}
\begin{proof} It suffices to show that
\begin{eqnarray}
d_{l, T} j^{(T-1)} = 0 , \label{complex1}\\
\delta^{(T-1)} d_{l, T} = 0.\label{complex2}
\end{eqnarray}
The equality $d_{l, T} j^{(T-1)} = 0$ holds, since $d_{l, T}$ is a quotient map of the linear operator $d_l$ acting
on ${\Omega}^*(M^{2n})$, and $j^{(T-1)}$ associates a cycle in $D^{p-(T-1), q+(T-1)}_l\subset H^{p+q}_l (M^{2n})$ to its class
in $E^{p,q}_{l, T}$.
The equality $\delta^{(T-1)} d_{l, T} = 0$ holds, since $\delta ^{(T-1)}d_{l, T} =\delta ^{(T-1)} j^{(T-1)} \delta ^{(T-1)} = 0$. This completes the proof of Lemma \ref{complex}.
\end{proof}
Let us continue the proof of Theorem \ref{nil}. From Lemma \ref{complex} we obtain the following long exact sequence
of the associated cohomology groups
\begin{equation}
(D^{p-(T-1), q+(T-1)}_l)^{(T-1)}\stackrel{j^*}{\to} E^{p,q}_{l, T+1 }\stackrel{\delta^*}{\to} (D^{p+1, q}_l) ^{(T-1)} \stackrel{{\partial}}{\to}
(D^{p+1, q}_l) ^{(T-1)} \label{vani3}
\end{equation}
\begin{lemma}\label{isom2} For $0\le p\le q\le n$ the connecting homomorphism ${\partial}: (D^{p+1, q}_l) ^{(T-1)}\to (D^{p+1, q}_l) ^{(T-1)}$ in (\ref{vani3}) is equal to the identity.
\end{lemma}
\begin{proof} By (\ref{vani2}) for any $x\in (D^{p+1, q}_l) ^{(T-1)}$ there exists $e\in E^{p,q}_{l,T}$ such that $y = \delta^{(T-1)} (e)$. Since $d^{p, q}_{l,T} (e) \in \ker \delta ^{(T-1)}$ there exists an element $y \in (D^{p+1, q}_l)^{(T-1)}$
such that $j^{(T-1)} (y) = d^{p,q}_{l, T}( e) = j ^{(T-1)} \delta^{(T-1)} (e)$. Since $j ^{(T-1)}$ is injective, $y = \delta ^{(T-1)} (e)$. By definition ${\partial} (x) = y = \delta ^{(T-1)} (e)$. It follows
that ${\partial} (\delta ^{(T-1)} e) = \delta ^{(T-1)} e$. This completes the proof of Lemma \ref{isom2}.
\end{proof}
\begin{corollary} \label{nilt} For $ p\ge T$ we have $E^{p,q}_{l, T +1} = 0$.
\end{corollary}
\begin{proof} For $p \ge T$ Lemma 5.11 yields the following exact sequence
$$(D^{*, *}_l)^{(T-1)}\stackrel{Id}{\to}(D^{*, *}_l)^{(T-1)}\stackrel{j^*}{\to} E^{p,q}_{l, T+1 }\stackrel{\delta^*}{\to} (D^{*, *}_l) ^{(T-1)} \stackrel{Id}{\to}
(D^{*, *}_l) ^{(T-1)}$$
which implies Corollary \ref{nilt} immediately.
\end{proof}
It follows from Corollary \ref{nilt} that $d_{l, T+1} : E^{p, q} _{l, T+1} \to E^{ p + T +1, q-T}_{l , T+1} =0$ for all $p \ge 0$. This completes the proof of Theorem \ref{nil}.
\end{proof}
We end this section with presenting a proof of the following stabilization theorem.
\begin{theorem}\label{stcom} (cf. \cite[Theorem 4]{PV2006}) Assume that $(M^{2n}, {\omega},\theta)$ is a compact connected globally
conformal symplectic manifold. Then the spectral sequence $(E^{p,q}_{l,k},d _{l,k})$ stabilizes at the $E^{*,*}
_{l,2}$-term.
\end{theorem}
\begin{proof} By Theorem \ref{conf} it suffices to prove Theorem \ref{stcom} for the case of a symplectic manifold $(M^{2n}, {\omega})$, i.e. $\theta = 0$.
The proof we present here uses many ideas in the proof of Theorem 2 in Di Pietro's Ph.D. Thesis \cite{Pietro2006} stated for connected compact symplectic manifolds.
By Lemma \ref{fk1} $E^{p, q}_{l, 1} = 0 = E^{p, q}_{l,k}$ if $q< p$ or $q>n$ for all $k\ge 1$. Thus it suffices to examine the terms $E^{p,p}_{l, k}$, $E^{p, p+r}_{l, k}$, for $0\le p\le n-r$, $r\ge 1$, $k\ge 2$.
\begin{lemma}\label{pp} Assume that ($M^{2n}, {\omega}, \theta)$ is a compact l.c.s. manifold, and $0\le p \le n$.\\
1. If $(M^{2n}, {\omega}, \theta)$ is a globally conformal symplectic manifold, then $E^{p, p}_{l, k} = {\mathbb R}$ for
all $l$ and $k\ge 2$. Moreover
$E^{p,p}_{l,k}$ is generated by the $p$-th power of the symplectic form ${\omega}$. \\
2. If $(M^{2n},{\omega}, \theta)$ is not conformal equivalent to a symplectic manifold,
then $E^{p,p}_{l, k} = 0$ for all $ p\not = l$ and for all $k \ge 1$.
\end{lemma}
\begin{proof}[Proof of Lemma \ref{pp}] By (\ref{0pl}) if $0 \le p \le n-1$ then
\begin{equation}
E^{p, p}_{l, 1} = H^{0}_{l-p} (M^{2n}).\label{i00}
\end{equation}
By (\ref{e13a}) we obtain
\begin{equation}
E^{n,n}_{l, 1} = C^0_{l-n} = C^\infty (M^{2n}).\label{nn1}
\end{equation}
We get from (\ref{nn1}) and (\ref{stab1a})
\begin{equation}
E^{n,n}_{l, 2} = C^\infty (M^{2n})/ d_{l,1} (E^{n-1, n}_{l,1}) = C^\infty (M^{2n})/ d^-_{l-n} ({\mathcal P}^1(M^{2n})) = H^{0}({\mathcal P}^*(M^{2n}), d^-_{l-n}), \label{nni}
\end{equation}
By Corollary \ref{lich0}
\begin{equation}
H_{0} ({\mathcal P}^*(M^{2n}), d^-_l) = H _{0}({\mathcal P}^*(M^{2n}), (d_{l-n})^*_{\omega}) =H^{2n}({\Omega}^*(M^{2n}), d_{l}).\label{nn2}
\end{equation}
Note that $d_{l,k} (E^{n,n}_{l,k}) = 0$ and ${\rm Im} \, d_{l,k} \cap E^{n,n}_{l,k} = 0$ for all $k \ge 2$. Using (\ref{nni}) we get
\begin{equation}
E^{n,n}_{l,k} = E^{n,n}_{l,2} = H_{0} ({\mathcal P}^*(M^{2n}), d^-_l) \text{ for all } k \ge 2. \label{nnil}
\end{equation}
Combining (\ref{nnil}), (\ref{nn2}) with Corollary \ref{lich0} we obtain the assertion of Lemma \ref{pp} for the cases $ p =0$ or $ p = n$.
Now let us consider $E^{p,p}_{k, r}$ with $0 < p < n$.
By (\ref{0pl}) $E^{p, p}_{l,1} = H^0 _{l-p} (M^{2n})$.
Let us first assume that $M^{2n}$ is globally conformal symplectic. Using Theorem
\ref{conf} we drop $l$ in the lower index of $d_{l, r}$ and $E^{p, q}_{l, r}$.
First we note that $E^{p,p}_0$ is generated by ${\omega}^p$. Since $E^{p,p}_k$ is a quotient of
$E^{p,p}_0$ and $[{\omega} ^p] \in E^{p, p}_k$ for all $k\ge 1$, taking into account $[{\omega}^n]\not = 0 \in E^{n,n}_k$, we complete the proof of Lemma \ref{pp}.1.
Now let us assume that $[\theta] \not = 0\in H^1 (M^{2n})$. Then
Corollary \ref{lich0} asserts that $H^0_l (M^{2n}) = 0$ for all $l\not = p$. It follows that $H^{p,p}_{l, \infty} = 0$
for all $l\not = p$. This complete the proof of Lemma \ref{pp}.
\end{proof}
\begin{lemma}\label{pp1} Assume that $(M^{2n}, {\omega})$ is a connected compact symplectic manifold. Then for $0\le p \le n-2$ and $k \ge 2$ we have $E^{p, p+1}_{ k} = H^1(M^{2n})$.
Furthermore $E^{n-1, n}_k = E^{n-1,n}_2$ for all $k \ge 2$.
\end{lemma}
\begin{proof}[Proof of Lemma \ref{pp1}] Using (\ref{stab1f}) and (\ref{stab1a}) we note that $d_1 : E^{0,1}_1 \to E^{1, 1}_1$ is equivalent to the map $d^-: H^{1} ({\mathcal P}^*(M^{2n}), d^+) \to H^0 (M^{2n},{\mathbb R}) = {\mathbb R}$. By (\ref{plus1b}) $H^{1} ({\mathcal P}^*(M^{2n}), d^+) = H^1 (M^{2n})$. Hence
\begin{equation}
E^{0,1}_2 = H^1(M^{2n}).\label{e012}
\end{equation}
It follows that, the image $d_k (E^{0,1}_k ) =0$ for all $k \ge 2$. Thus we get from (\ref{e012})
\begin{equation}
E^{0,1}_k = H^1 (M^{2n}) \text{ for all } k \ge 2.\label{01i}
\end{equation}
This proves Lemma \ref{pp1} for $E^{0,1}_k$, $k \ge 2$.
Since the operator $L^p :{\Omega}^ 1(M^{2n}) \to {\Omega} ^{2p +1}(M^{2n})$ is injective for all $p\le n-1$, using Lemma \ref{pp} we get
\begin{equation}
E^{0,1}_k \cong E^{0,1}_k \wedge E^{p,p}_k \subset E^{p, p+1}_k \text{ for all } k \ge 2.\label{pp1a}
\end{equation}
Note that $E^{p,p+1}_k $ is a quotient of $E^{p, p+1}_2$, which is isomorphic to $E^{0,1}_2$ by Proposition \ref{symm}. Taking into account (\ref{01i}) we obtain from (\ref{pp1a})
\begin{equation}
E^{p,p+1}_k \cong E^{0,1}_2 = H^1 (M^{2n}) \text{ for all } p \le n-2 \text { and } k \ge 2.\label{pp1b}
\end{equation}
This completes the proof the first assertion of Lemma \ref{pp1}. The second assertion of Lemma \ref{pp1} follows from the observation that $d_{2} (E^{n-1, n}_{2}) = 0 = {\rm Im}\, d_{2} \cap E^{n-1,n}_{l,2}$, and $d_k (E^{n-1,n}_k) = 0 = {\rm Im} \, d_k \cap E^{n-1, n} _k$ for all $ k \ge 3$.
\end{proof}
\begin{lemma}\label{pp2} Assume that $(M^{2n}, {\omega})$ is a connected compact symplectic manifold. Then $E^{n-2,n}_k = E^{n-2,n}_2$ for all $ k \ge 2$. Furthermore, for $0 \le p \le n -3$ and $k \ge 2$ we have
\begin{equation}
E^{p, p+2}_k \cong E^{0, 2}_2.\label{main1}
\end{equation}
\end{lemma}
\begin{proof} First we note that for $k \ge 2$
\begin{eqnarray}
d_k (E^{n-2,n}_k) = 0, \nonumber\\
{\rm Im}\, d_k \cap E^{n-2,n}_k = 0.\nonumber
\end{eqnarray}
Hence
\begin{equation}
E^{n-2,n}_k = E^{n-2,n}_2 \text{ for all } k \ge 2.
\end{equation}
This proves the first assertion of Lemma \ref{pp2}.
Next we observe that $d_2 (E^{0,2}_2 ) = 0$ and ${\rm Im} d_2 \cap E^{0,2}_ 2 = 0$.
Hence we get
\begin{equation}
E^{0,2}_k = E^{0,2}_2 \text{ for all } k \ge 2.\label{02k}
\end{equation}
Now we assume that $0\le p \le n-3$. Since $L^{p}: {\Omega} ^2 (M^{2n}) \to {\Omega} ^{2 + 2p}(M^{2n})$ is injective, using Lemma \ref{pp} we get
\begin{equation}
E^{0,2}_k \cong E^{0,2}_k \wedge E^{p, p}_k \subset E^{p, p+2}_k .\label{ppk}
\end{equation}
Since $E^{p,p+2}_k $ is a quotient group of $E^{p, p+2}_2$, which is isomorphic to $ E^{0,2}_2$ by Proposition \ref{symm}, using (\ref{02k}) and (\ref{ppk}) we get
\begin{equation}
E^{p,p+2}_k \cong E^{0,2}_2 \text{ for all } 0 \le p \le n -3.
\end{equation}
This completes the proof of Lemma \ref{pp2}.
\end{proof}
\begin{lemma}\label{pp3} We have $E^{p,p +r}_k = E^{0, r}_2 $ for all $k\ge 2$, $ p + r \le n-1$ and $r \ge 3$.
Furthermore $E^{n-r,n}_k = E^{n-r,n}_2$ for all $k \ge 2$ and $ r\ge 3$.
\end{lemma}
\begin{proof} We prove Lemma \ref{pp3} inductively on $r$ beginning with $r= 3$. For each $r$ we will consider $E^{p,p +r}_k $ with $k$ and $p$ increasing inductively.
First we note that
\begin{equation}
d_2 (E^{0,3}_2 ) = 0 \in E^{2,2}_2, \label{222}
\end{equation}
since $E^{2,2}_2 = E^{2,2}_k $ for all $ k \ge 2$ by Lemma \ref{pp}. From (\ref{222}) we obtain easily
\begin{equation}
E^{0,3}_k = E^{0,3}_2 \text { for all } k \ge 2.\label{222a}
\end{equation}
Now using the injectivity of the map $L^p : {\Omega} ^3(M^{2n}) \to {\Omega} ^{2p +3}(M^{2n})$ for $ p \le n -4$ and Lemma \ref{pp}, we get from (\ref{222a})
\begin{equation}
E^{0,3}_2 = E^{0,3}_k = E^{0,3}_k \wedge E^{p, p}_k \subset E^{p, p +3}_k .\label{222b}
\end{equation}
Since $E^{p, p+3}_k $ is a quotient group of $E^{p, p+3}_ 2 = E^{0,3}_2$, (\ref{222b}) implies
\begin{equation}
E^{p, p+3}_k = E^{0, 3}_2 \text { for all } 0\le p \le n-4, \, k \ge 2. \label{222c}
\end{equation}
This proves the first assertion of Lemma \ref{pp3} for $r=3$.
The second assertion of Lemma \ref{pp3} for $r =3$ follows from the identities ${\rm Im} d_k \cap E^{n-3,n}_k = 0 $ and $d_k ( E^{n-3,n}_k) = 0 \in E^{n +k-3,n-k +1}_k$ if $k \ge 2$, which is a consequence of Lemma
\ref{pp} if $k =2$.
Repeating this procedure we have for each $ n \ge r\ge 3$ the following sequences of identities with $k \ge 2$ and $ 0 \le p \le n-r$. First by induction on $r$ we get
\begin{equation}
d_2 (E^{0, r}_2) = 0 \in E^{2, r-1}_2,\label{fini1}
\end{equation}
since $E^{2, r-1}_2 = E^{2,r-1}_k$ for all $k \ge 2$ by the induction step. From (\ref{fini1}) we obtain immediately
\begin{equation}
E^{0, r}_k = E^{0, r}_2 \text { for all } k \ge 2\label{fini2}
\end{equation}
Since the map $L^p : {\Omega} ^r(M^{2n}) \to {\Omega} ^{2p +r}(M^{2n})$ for $ p \le n -r$ is injective, using Lemma \ref{pp}, we get from (\ref{fini2})
\begin{equation}
E^{0,r}_2 = E^{0,r}_k = E^{0,r}_k \wedge E^{p, p}_k \subset E^{p, p +r}_k .\label{fini3}
\end{equation}
Since $E^{p, p+r}_k $ is a quotient group of $E^{p, p+r}_ 2 $ which is isomorphic to $ E^{0,r}_2$ if $p +r \le n-1$ by Proposition \ref{symm}, (\ref{fini3}) implies
\begin{equation}
E^{p, p+r}_k = E^{0, r}_2 \text { for all } 0\le p \le n-r-1, \, k \ge 2. \label{fini4}
\end{equation}
Thus we get
\begin{equation}
E^{0, r}_k = E^{p, p+r}_k \text { for all } 0\le p \le n-r-1, \, k \ge 2.
\end{equation}
This completes the proof of the first assertion of Lemma \ref{pp3}.
The second assertion of Lemma \ref{pp3} for the inductive $r$ follows from the identities ${\rm Im} d_k \cap E^{n-r,n}_k = 0 $ and $d_k ( E^{n-3,n}_k) = 0 \in E^{n +k-r,n-k +1}_k$ if $k \ge 2$, which is a consequence of the induction assumption that $ E^{n+k-r, n-k+1}_k = E^{n+k-r, n-k+1}_2$ for $0\le k \le r-1$.
\end{proof}
Clearly Theorem \ref{stcom} follows from Lemmata \ref{pp}, \ref{pp1}, \ref{pp2} , \ref{pp3}.
\end{proof}
\begin{remark}\label{tsey} 1. Our
stabilization theorem \ref{stcom} gives an answer to the Tseng-Yau question on the relation between
the group $H^{p} (M^{2n}, d^+) = E^{0, p}_1 (M^{2n})$ for $ 0 \le p \le n-1$ and the cohomology groups $H^* (M^{2n}, {\mathbb R})$.
2. In the next section we show that if $(M^{2n}, {\omega})$ is a compact K\"ahler manifold, then
the spectral sequence stabilizes already at $E_1$-terms, see Theorem \ref{kaehler1}.
\end{remark}
\section{K\"ahler manifolds}
In this section we prove that
the spectral sequence $E^{p,q}_r$ stabilizes at the term $E^{*,*}_1$, if $(M^{2n},J, g)$ is a compact K\"ahler
manifold and ${\omega}$ is the associated symplectic form (Theorem \ref{kaehler1}).
Let $(M^{2n}, J, g, {\omega})$ be a compact K\"ahler manifold. As before, denote by $d^*$ the formal adjoint of $d$. Since the operator
$L$ commutes with the Laplacian $\triangle _d : = d d^* + d^* d$ we get the induced Lefschetz decomposition
of the space of harmonic forms on $M^{2n}$, and hence the induced Lefschetz decomposition of $H^*(M^{2n}, {\mathbb R})$. Let us denote by $P H^q(M^{2n}, {\mathbb R})$ the subset of primitive cohomology classes in $H^q(M^{2n}, {\mathbb R})$. Note that
each primitive cohomology class $[\alpha] \in P H^q (M^{2n})$ has a representative which is $\triangle_d$-harmonic and primitive.
\begin{proposition}\label{pmhk}(\cite[Proposition 3.18]{TY2009}) Assume that $(M^{2n}, J,g, {\omega})$ is a compact K\"ahler manifold. For $q \le n-1$ we have
$H_q ({\mathcal P}^*(M^{2n}), (d)^*_{\omega} ) = P H^q (M^{2n}) = H^{q} ({\mathcal P}^*(M^{2n}), d^+ )$.
\end{proposition}
\begin{proof} We give here another proof using \cite{Bouche1990}. Bouche proved that if $(M^{2n}, J, g, {\omega})$ is a compact K\"ahler manifold, the coeffective cohomology groups $ H^{2n-q} (\Cc^* M^{2n}, d)$ is isomorphic to the subgroup $H^{2n-q}_{\omega} ( M^{2n}) : = \{ x \in H^{2n-q }(M^{2n}, {\mathbb R})|\, Lx = 0\}$ for $0\le q \le n-1$ \cite[Proposition 3.1]{Bouche1990}. Next, using \cite[Corollary 2.4.2]{Brylinski1988}, or the following formula: ${\mathcal J} d ^* {\mathcal J} ^{-1} = d ^* _{\omega}$, which is proved in the similar way as (\ref{conj2}) replacing (\ref{formal10}) in the proof with (\ref{formal4}), we observe that the symplectic star operator $*_{\omega}$ maps $H ^{2n-q}_{\omega} (M^{2n})$ isomorphically onto the group $PH^q (M^{2n})$.
As we have noted
in Remark \ref{his1}.1, the coeffective cohomology group $H^q(\Cc^*M^{2n}, d)$ is isomorphic to the primitive homology group $H_{q} ({\mathcal P} ^*M^{2n})$. On the other hand Tseng-Yau proved that the group $H({\mathcal P}^*M^{2n}, (d)^-_{\omega})$ is isomorphic to the group $H_{q}({\mathcal P}^*(M^{2n}), d ^-)$ \cite[Lemma 2.7, part II]{TY2009} as well as to the group $H^{q,+} (M^{2n})$, \cite[Proposition 3.5, part II]{TY2009}, see also Proposition \ref{primi} and Proposition \ref{dpm1} above.
Combining these observations we complete the proof of Proposition \ref{pmhk}.
\end{proof}
\begin{theorem}\label{kaehler1} Assume that $(M^{2n}, J, g, {\omega})$ is a compact K\"ahler manifold. Then the spectral
sequence $E^{p, q}_r$ stabilizes at $E_1^{*,*}$.
\end{theorem}
\begin{proof} Since $(M^{2n}, J,g, {\omega})$ is a compact symplectic manifold, by Theorem \ref{stcom} the spectral sequence
$E^{p,q}_r$ stabilizes at $E_2$-terms. Thus to prove Theorem \ref{kaehler1} it suffices to show that
all the differentials $d_1: E^{p, q}_1 \to E^{p+1,q}_1$ vanish. By (\ref{stab1a}), if $q\le n-1$ then $d_{ 1} : E^{p,q}_{ 1} \to E^{p+1, q}_{1}$ is defined by the image of $d^- \tilde \alpha$, $\tilde \alpha \in {\mathcal P}^{q-p} (M^{2n})$. In this case it suffices to show that
any element $[\alpha]\in H^{q-p} ({\mathcal P}^*(M^{2n}), d^+)$ has a representative $\alpha \in {\mathcal P} ^{q-p}( M^{2n})$ such that
$d^- \alpha = 0$. By the Hodge theory for $d ^+ $ there is a harmonic representative $\alpha$ of $[\alpha]$ such that $(d^+) ^* \alpha = 0$. Lemma \ref{dpm} implies that for such harmonic form $\alpha$ we have
$d ^- \alpha = 0$. This implies $ d_1 (E^{p,q}_1) = 0$ for $q\le n-1$. It remains to consider
the image $d_1 (E^{p, n}_1)$. By (\ref{stab62}) it suffices to show that any element
$[\alpha] \in E^{p,n}_1$ has a representative $\alpha \in {\mathcal P}^{n-p}(M^{2n})$ such that $d^- (\alpha) =0$. Using the Hodge
theory for $d^+$ and (\ref{stab62}) we choose $\alpha$ to be the harmonic form. By Lemma \ref{dpm} $d ^- (\alpha) =0$. This completes the proof
of Theorem \ref{kaehler1}.
\end{proof}
\section{Examples}
In this section we consider two simple examples of compact l.c.s. manifolds and their primitive cohomologies. The first example is a nilmanifold of Heisenberg type \cite{Sawai2007b}, the second example is a 4-dimensional solvmanifold described in \cite{ACFM1989}, \cite{Banyaga2007}, \cite{Sawai2007}, \cite{HK2011}.
We calculate the primitive cohomology of these examples (Propositions \ref{mil1}, \ref{solv1}). We study some properties of primitive cohomology groups of a l.c.s. manifold, which is a mapping torus of a co-orientation preserving contactomorphism (Proposition \ref{inv2}). We show that the 4-dimensional solvmanifold is a mapping torus of a coorientation preserving contactomorphism of a connected contact 3-manifold, which is not isotopic to the identity (Theorem \ref{new1}).
\medskip
Let $H(n)$ denote the $(2n+1)$-dimensional Heisenberg Lie group and $\Gamma$ its lattice. It is well-known that the nilmanifold $N^{2n+2} : = (H(n)/\Gamma) \times S^1 $ has a canonical l.c.s. form ${\Omega}$, which we now describe following \cite{Sawai2007b}. Note that the Lie algebra ${\mathfrak h} (n) \oplus {\mathbb R}$ of $H(n) \times {\mathbb R}$
is given by $\langle X_i, Y_i, Z, A : [X_i, Y_i] = Z\rangle _{\mathbb R}$. We denote by $x_i, y_i, z, \alpha$ the dual basis.
Clearly $d\alpha = 0$ and $d{\Omega} = \alpha \wedge {\Omega}$. Here we use the same notations for the extension of $X_i, Y_i, Z, A, x_i, y_i, z, \alpha, {\Omega}$ to the right-invariants vector fields or differential forms on $H(n) \times {\mathbb R}$,
as well as for the descending vector fields or differential forms on $N^{2n+2}$.
\begin{proposition}\label{mil1} Let $(N^{2n+2}, {\Omega}, \alpha)$ be the l.c.s. nilmanifold described above. All the Lichnerowicz-Novikov cohomology groups $H^*({\Omega}^*(N^{2n+2}), d_{k\alpha})$ vanish, if $k \not = 0$. Consequently for $k \not = 0$ all the groups $E^{p,q}_{k, r}$, $r\ge 1$, of the associated spectral sequences vanish, unless $q = n$ and $r =1$. The group $E^{p,n}_{k,1}$ is infinite dimensional
for all $0 \le p \le n$.
\end{proposition}
\begin{proof} The first assertion of Proposition \ref{mil1} is a consequence of a result due to Millionshchikov, who proved that the Lichnerowicz-Novikov
cohomology groups $H^* ({\Omega}^*(M), d_\theta)$ of a compact nilmanifold $M$ always vanish unless $\theta$ presents a trivial cohomology class in $H^1 (M,{\mathbb R})$ \cite[Corollary 4.2]{Millionshchikov2005}.
The second assertion of Proposition \ref{mil1} for $E^{p,q}_{k, 1}$ is a consequence of the first assertion, combining with Lemma \ref{fk1} and the exact sequence (\ref{ex11}). Since ${\Omega} = d_\alpha (z)$, applying Theorem \ref{stab1} we obtain the second assertion from the first assertion combining with
the particular case $ r =1$ proved above. The third assertion follows from Lemma \ref{fk1} and from the ellipticity of the operators $d_k^+$, see the proof of Proposition \ref{dpm1}. This completes the proof of Proposition \ref{mil1}.
\end{proof}
Now we shall show an example of a l.c.s. 4-manifold $M_{k,n}$, which is an Inoue surface of type $S^-$, whose primitive cohomologies are non-trivial, and we will explain an implication of this non-triviality. The 4-manifold $M_{n,k}$ has been described in \cite{ACFM1989}, \cite{Banyaga2007}, \cite{Sawai2007}, \cite{HK2011}. Here we follow the exposition in \cite{Banyaga2007}.
Let $G_k$ be the group of matrices of the form
$$
\left(\begin{array}{clrr} e^{kz} & 0& 0& x \\
0 & e^{-kz} & 0 & y\\
0 & 0 & 1 & z\\
0 & 0 & 0 & 1
\end{array}\right),
$$
where $x,y,z \in {\mathbb R}$ and $k \in {\mathbb R}$ such that $e^k + e^{-k} \in \Z \setminus \{2\}$. The group $G_k$ is a connected solvable Lie group with a basis of right invariant 1-forms
\begin{equation}
dx -kxdz, \: dy + kydz, \: dz. \label{1inv}
\end{equation}
There exists a discrete subgroup $\Gamma _k \subset G_k$ such that $N_k = G_k/\Gamma_k$ is compact. The basis (\ref{1inv}) descends to a basis of 1-forms $\alpha, \beta, \gamma$ on $N_k$. The forms $\gamma$ and $\alpha \wedge \beta$ are closed and their cohomology classes generate $H^1 (M, {\mathbb R})$ and $H^2 (M, {\mathbb R})$ respectively.
Now let $\lambda \in {\mathbb R}$ be a number such that $\lambda [\alpha \wedge \beta] \in H^2 (M, \Z)$. For given $k,n$ denote by $M_{k,n}$ the total space of
the $S^1$-principal bundle over $N_k$ with the Chern class $n\lambda [\alpha \wedge \beta]$. Let $\eta$ be a connection form on $M_{k,n}$, equivalently
\begin{equation}
d\eta = n \lambda (\alpha \wedge \beta).\label{curv1}
\end{equation}
For simplicity we will denote the pull back to $M_{k,n}$ by the projection $M_{k,n} \to N_k$ of a form $\theta$ on $N_k$ again by $\theta$.
Banyaga showed that $M_{k,n}$ possesses many interesting l.c.s. structures. Here we consider only two l.c.s. forms $ d_{-k\gamma} \eta = n \lambda (\alpha \wedge \beta) - k \gamma \wedge \eta $ and $d_{k\gamma} \eta = n \lambda (\alpha \wedge \beta) + k \gamma \wedge \eta $ discovered by Banyaga \cite[Remark 2]{Banyaga2007}. Note that $M_{k,n}$ carries no symplectic structure, since $H^2 (M_{k,n}, {\mathbb R}) = 0$ \cite{ACFM1989}. Since
$M_{k,n}$ is compact, the Hodge theory applied to $d_{\pm \gamma}$ yields that $H^{2-i} ({\Omega} ^*(M_{n,k}), d_{\pm k \gamma}) = H^{2+i}({\Omega}^*(M_{n,k}), d_{\pm k \gamma})$. Since $[\pm k\gamma] \not = 0 \in H^1(M_{n,k}, {\mathbb R})$, the Lichnerowicz deformed
differential $d_{\pm k\lambda}$ is not gauge equivalent to the canonical differential $d$. Hence $H^0({\Omega} ^*(M^{2n}), d_{\pm k \gamma}) = 0$. Denote by ${\mathcal P}_{\pm}^*(M^{2n})$ the space of primitive forms corresponding to the
l.c.s. form $d_{\pm k\gamma} \eta$. Corollary \ref{lich0} yields that $H^{0} ({\mathcal P}^*_{\pm}(M_{k,n}), d^+_{\pm lk\gamma}) = 0$ for all $ l \not = 0$, and $H^0({\mathcal P}_{\pm} ^*(M^{2n}), d) = {\mathbb R}$.
\begin{proposition}\label{solv1} 1. $H^1 ({\Omega} ^*(M_{k,n}), d_{\pm k \gamma}) = {\mathbb R}$.\\
2. $H^2({\Omega} ^*(M_{k,n}), d_{\pm k \gamma} ) = {\mathbb R}$.\\
3. $ H^{1} ({\mathcal P}^*_{\pm} (M_{k,n}), d^+_{\pm k \gamma}) = {\mathbb R}^2$.
\end{proposition}
\begin{proof} It is known that $M_{k,n}$ is a complete solvmanifold. Indeed, the algebra ${\mathfrak g}_{k,n}$ of the corresponding solvable group possesses the basis $(X, Y, Z, T)$ dual to $(\alpha, \beta, \gamma, \eta)$ with the following properties \cite{ACFM1989}, or see (\ref{solv2d}) and (\ref{solv3d}) below.
\begin{eqnarray} [X, Z] = k X, \, [X, Y] = -n \lambda T, \, [Y, Z] = - k Y, \label{solv2}\\
\, [X, T] = [Y, T] = [Z, T] = 0.\label{solv3}
\end{eqnarray}
Using (\ref{solv2}) and (\ref{solv3}) we observe that the Lie subalgebras $\langle T\rangle_{\mathbb R} \subset \langle T, X \rangle _{\mathbb R} \subset \langle T, X, Y \rangle _{\mathbb R}$ are ideals of ${\mathfrak g}_{k,n}$, so $M_{k,n}$ is completely solvable.
Now we apply the result
by Millionshchikov \cite[Corollary 4.1, Theorem 4.5]{Millionshchikov2005}, which reduces the computation of the Novikov
cohomology groups of a compact complete solvmanifold $G/\Gamma$ to the computation of the induced Novikov cohomology groups of the Lie algebra ${\mathfrak g}$ of $G$. For our computation it is useful to rewrite (\ref{solv2}) and (\ref{solv3}) in the dual basis of ${\mathfrak g}^*_{k,n}$, or using the explicit formulae for $\alpha, \beta, \gamma, \eta$ given in (\ref{1inv}), (\ref{curv1}) above to obtain
\begin{eqnarray}
d\alpha =- k \alpha \wedge \gamma, \: d\beta = k \beta \wedge \gamma, \label{solv2d}\\
d\gamma = 0, \: d\eta = n \lambda (\alpha \wedge \beta). \label{solv3d}
\end{eqnarray}
1. Abbreviate $d_{\pm k \gamma}$ as $d_{\pm k}$. Using (\ref{solv2d}) and (\ref{solv3d}) we get
\begin{eqnarray}
d_{\pm k} \alpha = (k \pm k) \gamma \wedge \alpha, \: d_{\pm k} \beta = (- k \pm k) \gamma \wedge \beta , \label{solv4d}\\
\gamma = d_{\pm k } (\pm 1/k), \: d_{\pm} \eta = n \lambda (\alpha \wedge \beta) \pm k\gamma \wedge \eta.\label{solv5d}
\end{eqnarray}
Using (\ref{solv4d}) and (\ref{solv5d}) it is easy to compute that $\alpha$ is a generator of $H^1 ({\Omega}^*(M_{k,n}), d_{-k})$, and $\beta$ is a generator
of $H^1 ({\Omega}^*(M_{k,n}), d_k)$. This proves the first assertion of Proposition \ref{solv1}.
2. For computing $H^2 ({\Omega} ^*(M_{k,n}), d_{\pm k})$ we use (\ref{solv4d}), (\ref{solv5d}), and the following formulae
\begin{eqnarray}
d_{\pm k} (\alpha \wedge \beta ) = \pm k \alpha \wedge \beta \wedge \gamma, \nonumber\\
d_{-k} (\alpha \wedge \gamma)= 0, \: d_{k} (\beta \wedge \gamma ) = 0, \nonumber\\
d_{\pm k} (\alpha \wedge \eta) = (-k \pm k) \gamma \wedge \alpha \wedge \eta, \: d_{\pm k} (\beta \wedge \eta) = (k \pm k)\gamma \wedge \beta \wedge \eta, \nonumber \\
d_{\pm k} (\gamma \wedge \eta) = - n \lambda \alpha \wedge \beta \wedge \gamma.\nonumber
\end{eqnarray}
It is easy to see that $\alpha \wedge \eta$ is a generator of $H^2 ({\Omega} ^*(M_{k,n}), d_{-k})$ and $\beta \wedge \eta$
is a generator of $H^2({\Omega}^*(M_{k,n}), d_k) $. This proves the second assertions of Proposition \ref{solv1}.
3. The third assertion of Proposition \ref{solv1} is a consequence of the first assertion and Formula(\ref{plus1c}).
This completes the proof of Proposition \ref{solv1}.
\end{proof}
In the remaining part of this section we study some properties of primitive cohomology groups of l.c.s. manifolds associated with a co-orientation preserving contactomorphism. We show that the l.c.s. solvmanifold studied before is an example of a l.c.s. manifold associated with a non-trivial contactomorphism.
Let $(M^{2n+1}, \alpha)$ be a co-orientable contact manifold and $f $ be a co-orientation preserving contactomorphism of $(M^{2n+1},\alpha)$, i.e. $f^* (\alpha) = e ^h \cdot \alpha$ for some
$h \in C^\infty (M^{2n})$.
The mapping torus $M_f^{2n+2} = (M\times [0,1])/( [x, 0] = [f (x) , 1])$ of a contactomorphism $f$ is a fibration over $S^1$ whose fiber is $M^{2n+1}$. Let us denote this fibration by $\pi: M_f^{2n+2} \to S^1$ with
$\pi ^{-1} (s) = [M, s]$.
Let $f_t: M_f^{2n +2} \to M_f^{2n +2}$ be a 1-parameter family of diffeomorphisms defined by:
$$f_t ([x, s]) = [x, s+t \mod 1] \text { for } t \in {\mathbb R}.$$
In particular $f_1 ([x, 0]) = [f(x), 0])$. Let us also denote by $\alpha$ the contact 1-form on $[M, 0]$ obtained by identifying $M$ with $[M, 0]$.
Let $B$ be the vector field on $M^{2n+2}_f$ defined by $B([x,s]) = (d/dt)_{|t= 0} f_t ([x, s])$. Since $f ^* (\alpha) = e ^h \alpha$ the following 1-form $\tilde \alpha$
\begin{equation}
\tilde \alpha (x, t)_{|\pi^{-1} (t)}: = e ^{ - th (x)}f_{-t} ^ * (\alpha), \, \tilde \alpha (B) = 0.\label{ext1}
\end{equation}
is well-defined on $M^{2n+2}_f$, moreover
$$ f _t ^* ( \tilde \alpha) _{| \pi ^{-1} (t)} = \tilde \alpha _{|\pi ^{-1} (0)} \text { for all } 0\le t \le 1. $$
Set $\theta : = \pi ^* (dt)$.
\begin{proposition}\label{mtc} (cf. \cite[Proposition 3.3.]{BK2010}) 1. Assume that $ (M^{2n +1}, \alpha)$ is a compact co-orientable contact manifold and $f$ is a co-orientation preserving contactomorphism. There exists a positive number $c_0$ such that $(M^{2n+2}_f, {\omega}_c:=d\tilde \alpha + c\theta \wedge \tilde\alpha, c\theta)$ is a l.c.s. manifold for
all $ c\ge c_0$.
2. Assume that $f$ preserves the contact 1-form $\alpha$. Then $(M^{2n+2}_f, {\omega}:=d\tilde \alpha + \theta \wedge \tilde\alpha, \theta)$ is a l.c.s. manifold.
\end{proposition}
\begin{proof} 1. Clearly (\ref{ext1}) implies that $ rk \, d\tilde \alpha \ge rk \, d\alpha = 2n$. Using this we conclude that there exists a positive number $c_0$ such that $rk\, d{\omega}_c = 2n +2$ for all $ c\ge c_0$, since $M^{2n +1}$ is compact. Further, $d (c \theta ) = 0$ and
$ {\omega}_c = d _{c\theta} (\tilde \alpha)$. This proves that $(M^{2n +2} _f, {\omega} _c, c\theta)$ is a l.c.s. manifold.
2. Assume that $f ^* (\alpha )= \alpha$. Then $\tilde \alpha ( [x, t])_{\pi ^{-1} (t)} = f ^* _{-t} \alpha$. It follows that $rk\, d\tilde \alpha = rk\, d\alpha = 2n$, and $ rk \, {\omega} _1 = 2n +2$.
Hence ${\omega} _1 = {\omega}$ is a l.c.s. form, taking into account ${\omega} = d_{\theta } \alpha$.
\end{proof}
\begin{proposition} \label{inv2} 1. Suppose that $f_0$ and $f_1$ are co-orientation preserving contactomorphisms of a compact co-orientable contact manifold $(M^{2n+1}, \alpha)$. The l.c.s. manifolds $M^{2n+2}_{f_0}$ and
$M^{2n +2}_{f_1}$ are diffeomorphic, if $f_0$ and $f_1$ are isotopic. For sufficiently large number $c$ the primitive cohomology groups of $(M^{2n +2}_{ f_0}, {\omega} _c, c \theta)$ and of $(M^{2n +2}_{ f_1}, {\omega} _c', c \theta)$ are isomorphic.
2. Let $\theta$ be the Lee form of the associated l.c.s form on $M^{2n +1}_f$. If $f$ is isotopic to the identity, the Lichnerowicz cohomology groups $H ^*({\Omega} ^* (M^{2n+2}_f), d_{c \theta})$ are zero, for any $c \not = 0$.
\end{proposition}
\begin{proof} The first assertion of Proposition \ref{inv2}.1 is well-known.
The second assertion of Proposition \ref{inv2}.1 is a consequence of Theorem \ref{conf}, observing that ${\omega} _c - {\omega} '_c = d_{c\theta} (\tilde \alpha - \tilde \alpha ')$.
Finally Proposition \ref{inv2}.2 follows from the first assertion, combining with the fact that the l.c.s. manifold $(M^{2n+1} \times S^1, d_\theta \tilde\alpha, \theta)$ associated to the identity mapping of the contact manifold $(M^{2n+1}, \alpha)$ has
vanishing Lichnerowicz-Novikov groups, taking into account the K\"unneth formula and the formula $H^* ({\Omega} ^*(S^1), d_{c dt}) = 0$ if $c \not = 0$.
This completes the proof of Proposition \ref{inv2}.
\end{proof}
Now we shall show that our l.c.s. manifold $(M_{k,n}, d_{k\gamma}\eta, k\gamma)$ is a mapping torus of a non-trivial co-orientation preserving contactomorphism.
First we prove the following
\begin{proposition}\label{cri1} Assume that $\gamma$ is a closed 1-form on a compact smooth manifold $M$. If $[\gamma]\in H^1 (M, \Z)$ and $\gamma$ is now-where vanishing, then there
is a submersion $f : M \to S^1$ such that $f^* (dt) = \gamma$, where $dt$ is the canonical 1-form on $S^1$.
\end{proposition}
\begin{proof} We use Tischler's argument in \cite{Tischler1970}.
Since $S^1$ is the Eilenberg-Maclane space there exists a map $f_1: M \to S^1$ such that $f^*([dt]) = [\gamma]$. Without loss of generality we assume that
$f$ is a smooth map. Hence we have $f^* (dt) = \gamma + dh$ for some smooth function $h$ on $M$. Now we observe that $f_1 ^* (dt) + dh = (f_1 + \Pi \circ h) (dt)$, where $\Pi : {\mathbb R} \to S^1$ is the natural projection. Clearly
the map $f = f_1 + \Pi\circ h$ is a submersion, since $\gamma$ is no-where vanishing. This completes the proof of Proposition \ref{cri1}.
\end{proof}
Now we are ready to show the following implication of Proposition \ref{solv1}.
\begin{theorem}\label{new1} The l.c.s. manifold $(M_{k,n}, d_{k\gamma}\eta, k\gamma)$ is a mapping torus
of a coorientation preserving contactomorphism $f$ of a 3-dimensional connected contact manifold. Moreover $f$ is not isotopic to the identity.
\end{theorem}
\begin{proof} Since $H^1 (M_{k,n}, {\mathbb R}) = {\mathbb R}$ \cite{ACFM1989}, and $d\gamma = 0$, there exists a positive number $p$ such that $p [\gamma]$ is a generator of $H^1 (M_{k,n}, \Z) = \Z = Hom (H_1(M_{k,n}, \Z), \Z)$ \cite[Chapter VI, 7.22]{Dold1972}. Applying Proposition \ref{cri1} we conclude that
$M_{k,n}$ is a fibration over $S^1$ whose fibers are the foliation ${\mathcal F}_1 : =\{\gamma = 0\}$, and $f ^* (dt) = p \cdot\gamma$, since $\gamma$ is nowhere vanishing. Denote by $\pi: M_{k,n} \to S^1$ the corresponding fibration.
Note that the restriction of $\eta$ to each fiber $\pi^{-1}(t)$, $ t \in S^1$, is a contact form, since $X, Y, T$ are tangent to the fiber and we have $\eta (T) = 1$, $d\eta (X, Y)\not = 0$.
First we will show that the fiber $F : = \pi^{-1}(t)$, $t \in S^1$, is connected.
Let us consider the following exact sequence of homotopy groups
\begin{equation}
\pi_1(M_{k,n}) \to \pi_1 (S^1) \to \pi_0 (F) \to 0 = \pi _0 (M_{k,n}).\label{homot1}
\end{equation}
To show that $ \pi_0 (F) = 0$ it suffices to prove that the map $\pi_1 (M_{k,n}) \to \pi_1 (S^1)$ is surjective. Since $p [\gamma]$ is a generator of $H^1 (M_{k,n}, \Z)$ there exists an element $a \in H_1 (M_{k,n}, \Z)$ such that
$ \langle p[\gamma] , a\rangle = 1$. Since $\langle [dt], \pi_* (a) \rangle = \langle p[\gamma] , a\rangle = 1$, it follows that $\pi_*: H_1 (M_{k,n}, \Z) \to H_1 (S^1)$ is surjective. Hence $\pi_*: \pi_1 (M_{k,n}) \to \pi_1 (S^1)$ is
surjective. Hence $F$ is connected.
Now let $f_t$ denote the flow on $M_{k,n}$ generated by the vector field $Z$. We note that $\Ll_Z (\gamma) = d (\gamma (Z)) = 0$, so $f_t$ respects fibration $\pi$. Next
we have $\Ll_Z ( \eta) = Z \rfloor n\lambda \alpha \wedge \beta + d (\eta (Z)) = 0$. Hence $f_t$ preserves also the contact form on the fiber $F$. This proves the first assertion.
The second assertion
is a consequence of Proposition \ref{solv1} and Proposition \ref{inv2}. This completes the proof of Theorem \ref{new1}.
\end{proof}
\section* {Acknowledgement}
H.V.L. thanks Alexandre Vinogradov for explaining the idea of their paper \cite{PV2006} during the conference on integrable systems at Hradec nad Moravici in October 2010. We acknowledge Alexandre Vinogradov kindness for lending H.V.L. the Ph.D. Thesis of Di Pietro in early summer 2011, which accelerated our work over sections 4 and 5 considerably. H.V.L. also thanks Thiery Bouche for sending his reprint \cite{Bouche1990}, Petr Somberg for informing her of the paper by Rumin \cite{Rumin1994}, J\"urgen Jost for his support and his invitation to give a lecture on this subject at the Max-Planck-Institute in Leipzig in March 2011, and the ASSMS, Government College University, Lahore-Pakistan for their hospitality and financial support during her visits in February and September 2011, where a part of this note has been written.
Last but not least, H.V.L. would like to express her gratitude to Kaoru Ono and Lorenz Schwachhofer for their helpful remarks.
\medskip
|
\section{\bf Introduction}
The paradigm of primordial inflation is, by far, the most satisfactory explanation
for early universe phenomena \cite{rocher}. As a general prescription, inflation occurs
due to a slowly rolling scalar field, the inflaton,
dynamically giving rise to an epoch of accelerated
expansion dominated by a false vacuum \cite{barrow}.
Primordial quantum fluctuations of inflaton are responsible for creation of matter
content and observed perturbations in the Cosmic Microwave Background Radiation (CMBR).
Further, slow-roll inflationary scenario generically predicts almost Gaussian adiabatic perturbations with a nearly flat
spectrum, which conforms well with the latest observations.
Recently, some interesting proposition of inflationary model building was brought forth by Minimally Supersymmetric Standard Model (MSSM)
where the inflaton is a gauge invariant \cite{enqvist,arindam}
$n=4$ level combination of scalar superpartners squark and slepton
fields and fermionic superpartner gauginos which are
candidate Cold Dark Matter (CDM) particles. However the original potential for $n=4$ level
is unable to extract a suitable symmetry along the flat direction. To serve this purpose
the usual way is to incorporate {\it saddle point mechanism} to the
MSSM potential leading to vanishing of the second derivative
and the slow roll phase is driven by the next leading order derivative
of the potential \cite{rocher,enqvist,arindam,maz,maz1}.
In most of the phenomenological situations, a fine tuning mechanism
is needed to place the flat direction field to the immediate neighborhood of the {\it saddle point}.
It is worthwhile to mention that MSSM inflation occurs at a comparative
lower scale. This is in strong contrast with the
conventional class of models where the unfamiliar inflaton couplings to Standard Model (SM)
are originated through arbitrary gauge singlets
leading to the field magnitudes at GUT scale or higher, and hence,
face problems in satisfactory quantitative estimation of
a huge sector of the post-inflationary evolution i.e. thermal history
of the early Universe, baryon
asymmetry and CDM. Herein lies the most appealing feature of MSSM inflation for
which known SM couplings are measurable in
laboratory experiments such as Large Hadron Collider (LHC) \cite{lhc} or
future linear colliders.
In the present article we will consider a specific MSSM scenario where, for a specific choice of
soft supersymmetry (SUSY) breaking parameters A (trilinear couplings) and the inflaton mass $m_{\phi}$,
the potential is D-flat along the
$\textbf{QQQL,QuQd,QuLe}$ and $\textbf{uude}$
directions. For our model existence of {\it saddle point} is guaranteed by the non-vanishing fourth derivative
of the potential, which makes the potential more flat than the previous ones.
This implies more precise information in the RG flow.
As we will show, this is the highest level of precision constraint one can impose
on RG flow keeping the effective potential renormalizable
in the vicinity of the {\it saddle point}. Our primary intention is to investigate for the analytical as well
as the numerical expressions for different observational parameters
for MSSM inflation with these new flat directions. As it will turn out, they match quite well with latest observational
data from WMAP7 \cite{wmap07} and are expected to fit well with
upcoming data from PLANCK \cite{planck}. Additionally we have explicitly shown the connection between
running and running of the running of spectral index to the Primordial Black Hole (PBH) formation.
To this end we get the fine tuned parameter space
which is also in
good agreement with present estimates of cosmological frameworks.
We have further explored features of the MSSM from the solution of one
loop RGE which could be measured by LHC or future linear collider.
\section{\bf Flat directions and potential around saddle point }
Let us start with $n=4$ level superpotential \cite{martin}
\begin{equation}\begin{array}{ll}\label{mssm}
W^{nr}_{4} = \frac{1}{M}\left[\sum_{I=1}^{24}\alpha_I ({\bf QQQL})_I+
\sum_{I=1}^{81}\beta_I ({\bf QuQd})_I+\right.\\ \left.
~~~~~~~~~~~~\sum_{I=1}^{81}\gamma_I ({\bf QuLe})_I
+\sum_{I=1}^{27}\delta_I ({\bf uude})_I\right]
;\end{array}\end{equation}
The renormalizable flat directions of the MSSM at n=4 level correspond to the gauge invariant
monomials subject to the four additional complex constraints \cite{martin} two each from
\begin{equation} F^{\alpha}_{H_{u}}=\mu H^{\alpha}_{d}+\lambda^{ab}_{U}Q^{\alpha}_{a}u_{b}=0,\end{equation}
\begin{equation} F^{\alpha}_{H_{d}}=-\mu H^{\alpha}_{u}+\lambda^{ab}_{D}Q^{\alpha}_{a}d_{b}+\lambda^{ab}_{E}L^{\alpha}_{a}e_{b}=0,\end{equation}
which can lift the flat directions which do not contain a Higgs field. Here $\lambda_{U}, \lambda_{D}$ and $\lambda_{E}$ are the Yukawa
couplings, $H_{u}, H_{d}$ are the Higgs superfield and the $\mu$- term appears in the renormalizable part of the superpotential of MSSM . Consequently the equation(\ref{mssm})
breaks into four parts, each one of them now being flat:
\begin{eqnarray}\label{xd1}W^{(1)}_{4}&=&
\frac{1}{M}\sum_{I=1}^{24}\alpha_I ({\bf QQQL})_I,
\\
\label{xd2} W^{(2)}_{4}&=&
\frac{1}{M}\sum_{I=1}^{81}\beta_I ({\bf QuQd})_I,
\\
\label{xd3} W^{(3)}_{4}&=&
\frac{1}{M}\sum_{I=1}^{81}\gamma_I ({\bf QuLe})_I,
\\
\label{xd4} W^{(4)}_{4}&=&
\frac{1}{M}\sum_{I=1}^{27}\delta_I ({\bf uude})_I,\end{eqnarray}
resulting in $W^{(i)}_{4}\approx \frac{\lambda_{4}}{4M}{\bf\Phi}^{4} ~\forall i(=(1,2,3,4))$.
Considering any one of the above flat directions leads to the one loop corrected effective potential
\begin{equation}\label{vg}
V(\phi,\theta)=\frac{1}{2}m^{2}_{\phi}\phi^{2}
+\frac{\lambda_{4}A}{4M}\phi^{4}Cos(4\theta+\theta_{A})
+\frac{\lambda^{2}_{4}}{M^{2}}\phi^{6},\end{equation}
for all i. Here we define $\lambda_{4}=\lambda_{4,0}\left[1+D_{3}\log\left(\frac{\phi^{2}}{\mu^{2}_{0}}\right)\right]$,
$A=\frac{A_{0}\left[1+D_{2}\log\left(\frac{\phi^{2}}{\mu^{2}_{0}
}\right)\right]}{\left[1+D_{3}\log\left(\frac{\phi^{2}}{\mu^{2}_{0}}\right)\right]}$ and
$m^{2}_{\phi}=m^{2}_{0}\left[1+D_{1}\log\left(\frac{\phi^{2}}{\mu^{2}_{0}}\right)\right]$
and in $\bf G_{MSSM}=SU(3)_{C}\otimes SU(2)_{L}\otimes U(1)_{Y}$ the representative flat direction field content is given by
\begin{equation}\begin{array}{llll}\label{uip}
{\bf Q^{I_{1}}_{a}}=\frac{1}{\sqrt{2}}({\bf\Phi},0)^{T},{\bf Q^{I_{2}}_{b}}=\frac{1}{\sqrt{2}}({\bf\Phi},0)^{T},{\bf Q^{I_{3}}_{c}}=\frac{1}{\sqrt{2}}({\bf\Phi},0)^{T},\\
~~~~~~~~ {\bf L^{I_{4}}_{3}}=\frac{1}{\sqrt{2}}P_{d}(0,{\bf\Phi})^{T},{\bf d^{B_{1}}_{a}}=\frac{{\bf\Phi}}{\sqrt{2}},
{\bf u^{B_{2}}_{b}}=\frac{{\bf\Phi}}{\sqrt{2}},\\~~~~~~~~~~~~~~~~~~~~~~ {\bf u^{B_{3}}_{c}}=\frac{{\bf\Phi}}{\sqrt{2}},{\bf e_{3}}=\frac{{\bf\Phi}}{\sqrt{2}}.
\end{array}\end{equation}
Here $m_{0},A_{0}$ and $\lambda_{4,0}$ are the values of the respective parameters at the scale $\mu_{0}$ and $D_{1},D_{2}$
and $D_{3}$ ($|D_{i}|\ll 1 \forall i$) are the fine tuning parameters. Additionally in the field contents
${\bf 1 \leq B_{1},B_{2},B_{3} \leq 3}$ are color indices, ${\bf 1 \leq a,b,c\leq 3 }$ denote the indices for quark and lepton families and
${\bf 1\leq I_{1},I_{2},I_{3},I_{4}\leq 2}$ are the weak isospin indices. The
flatness constraints require that ${\bf B_{1} \neq B_{2} \neq B_{3}}$ for quarks, ${\bf I_{1}\neq I_{2}\neq I_{3}\neq I_{4}}$,
${\bf\sum^{3}_{d=1}P^{2}_{d}=1}$ ${\bf\forall P_{d}\in \mathbb{R}}$ for leptons and ${\bf a\neq b\neq c}$ for both. In eqn(\ref{vg}) $m_{\phi}$
represents the soft SUSY breaking mass term, $\phi$ the radial coordinate of the complex scalar field
${\bf\Phi}=\phi\exp(i\theta)$ ($\in\mathbb{C}$) and the second term is the so called
A-term which has a periodicity of $2\pi$ in 2 D along with
an extra phase $\theta_{A}$. The radiative correction slightly affects the soft term and the value
of the saddle point.
For $n=4$ we get an extremum for the principal values of $\theta$ at $\theta=\frac{(m\pi-\theta_{A})}{4}$
(where ${\bf m\in \mathbb{Z}}$)
\begin{equation}\begin{array}{ll}\label{sd} \phi_{0}=\sqrt{\frac{M}{4\lambda_{4}(3+D_{3})}}\left[A\left(1+\frac{D_{2}}{2}\right)\right.
\\ \left.\pm \sqrt{A^{2}\left(1+\frac{D_{2}}{2}\right)^{2}-8m^{2}_{\phi}(1+D_{1})(3+D_{3})}\right]^{\frac{1}{2}},
\end{array}
\end{equation}
which appears from the constraint $V^{'}(\phi_{0})=0$ as a
necessary condition for {\it saddle point}. However, this condition alone will not lead to
{\it saddle point}. Rather, we have to make the potential sufficiently flat which can be achieved by vanishing
higher derivatives of the potential. In this article, we consider non-vanishing fourth derivative of the potential
resulting in saddle point. This will imply more fine-tuning but increased precision level in the information obtained
from RG flow. Below we demonstrate how this is materialized.
As discussed, $V^{''''}(\phi_{0})<0$ will give us secondary local minimum.
This leads to constraint relations:
\begin{equation}\label{Ast}A =\sqrt{2(3+D_{3})G_{1}G_{2}G_{3}}m_{\phi}(\phi_{0}),\end{equation}
\begin{equation}\begin{array}{llll} \label{con2} D_{3}=\frac{MA_{0}}{4\lambda_{4,0}\phi^{2}_{0}\left(37+60\log\left(\frac{\phi_{0}}{\mu_{0}}\right)\right)}
\left\{D_{2}\left(13+12\log\left(\frac{\phi_{0}}{\mu_{0}}\right)\right)\right.\\
\left.~~~~~~~~~~~~~~~~~~~~~~~~~~-\frac{2m^{2}_{\phi}(\phi_{0})D_{1}M}{\lambda_{4,0}A_{0}\phi^{2}_{0}}
+6\left(1-\frac{20\lambda_{4,0}\phi^{2}_{0}}{MA_{0}}\right)\right\}\end{array},\end{equation}
one each for $V^{''}(\phi_{0})=0$ and $V^{'''}(\phi_{0})=0$. In this context $G_{1}=\left[\frac{(1+D_{1})}{(3+D_{3})}(15+11D_{3})-(1+3D_{1})\right]^{2}$,
$G_{2}=\left[(1+D_{1})\left(3+\frac{7}{2}D_{3}\right)-(1+3D_{1})\left(1+\frac{D_{2}}{2}\right)\right]^{-1}$,
$G_{3}=\left[\frac{\left(1+\frac{D_{2}}{2}\right)}{(3+D_{3})}(15+11D_{3})-\left(3+\frac{7}{2}D_{3}\right)\right]^{-1}$.
For the limit $|D_{1}|\ll 1$,$|D_{2}|\ll 1$ and $|D_{3}|\ll 1$ which gives
$\phi_{0}=\phi^{tree}_{0}\left[1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right]^{\frac{1}{2}}$ and
$A\simeq A_{tree}\left[1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right]$,
where $\phi^{tree}_{0}=\sqrt{\frac{m_{\phi}(\phi_{0})M}{\lambda_{4}\sqrt{6}}}$ and $A_{tree}=2\sqrt{6}m_{\phi}(\phi_{0})$
represents tree level expressions. This means, during RG flow mentioning two parameters only ($D_{1}$ and $D_{2}$) will
suffice instead of the usual three parameters in earlier MSSM models. This results in more precise information in RG flow.
One may get tempted to vanish further higher derivatives of the potential in order to evaluate other
unknown parameters ($D_{1}$ and $D_{2}$) without going into RG flow but this will make the effective inflaton potential
in the vicinity of {\it saddle point} non-renormalizable. So, this is the highest level of precision
constraint one can impose on RG flow parameters.
Consequently, around the {\it saddle point} $\phi_{0}$, the inflaton potential can be expanded in a Taylor series as,
\begin{equation}\label{hgkl} V(\phi)=\tilde{C}_{0}+\tilde{C}_{4}(\phi-\phi_{0})^{4},\end{equation}
where $\tilde{C}_{0}=V(\phi_{0})=\frac{m^{3}_{\phi}(\phi_{0})M}{6\sqrt{6}\lambda_{4}}
\left\{3\left(1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right)\right.\\ \left.\left[1+D_{1}\log\left(\frac{\phi^{2}_{0}}{\mu^{2}_{0}}\right)\right]-3\left(1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right)^{2}\left[1+D_{2}\log\left(\frac{\phi^{2}_{0}}{\mu^{2}_{0}}\right)\right]\right.
\\ \left.+\left(1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right)^{2}\left[1+D_{2}\log\left(\frac{\phi^{2}_{0}}{\mu^{2}_{0}}\right)\right]
\right\}$
and\\
$\tilde{C}_{4}=\frac{1}{{4!}}V^{''''}(\phi_{0})
=\frac{m^{2}_{\phi}(\phi_{0})}{24\sqrt{6}\phi^{2}_{0}}\left(1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right)
\\ \left\{\left\{\left[\left(\frac{360}{\sqrt{6}}-12\sqrt{6}\right)+(684D_{3}-50\sqrt{6}D_{2})\right]
\left(1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right)\right.\right.
\\ \left.\left.-\frac{2\sqrt{6}D_{1}}{\left(1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right)}\right\}
+\left(1+\frac{D_{1}}{2}-\frac{D_{3}}{6}\right)\left(\frac{360D_{3}}{\sqrt{6}}
-12\sqrt{6}D_{2}\right)\right.\\ \left.\log\left(\frac{\phi^{2}_{0}}{\mu^{2}_{0}}\right)\right\}$.
In what follows we shall model MSSM inflation with the above potential.
\section{\bf Modeling MSSM inflation and parameter estimation}
For brevity, let us introduce a change of parameter $\phi$ $\rightarrow$ $x=\phi-\phi_{0}$
which represents the inflaton with shifted origin.
Using this new notation of field the slow roll parameters \cite{lyth} are given by,
\begin{eqnarray}\label{df1}\tiny\epsilon_{v}(x)&=&
\frac{M^{2}}{2}\left(\frac{V^{'}}{V}\right)^{2}
=\frac{8\tilde{C}^{2}_{4}M^{2}x^{6}}
{(\tilde{C}_{0}+\tilde{C}_{4}x^{4})^{2}},
\\
\label{df2} \tiny\eta_{v}(x)&=&
M^{2}\left(\frac{V^{''}}{V}\right)
=\frac{12\tilde{C}_{4}M^{2}x^{2}}
{(\tilde{C}_{0}+\tilde{C}_{4}x^{4})},
\\
\label{df3}\tiny\xi^{2}_{v}(x)&=&
M^{4}\left(\frac{V^{'}V^{'''}}{V^{2}}\right)
=\frac{96\tilde{C}^{2}_{4}M^{4}x^{4}}
{(\tilde{C}_{0}+\tilde{C}_{4}x^{4})^{2}},
\\
\label{df5}\tiny\sigma^{3}_{v}(x)&=&
M^{6}\left(\frac{(V^{'})^{2}V^{''''}}{V^{3}}\right)
=\frac{384\tilde{C}^{3}_{4}M^{6}x^{6}}{(\tilde{C}_{0}+\tilde{C}_{4}x^{4})^{3}},\end{eqnarray}
where a prime denotes $d/d\phi=d/dx$.
During slow-roll inflation $\epsilon_{v},|\eta_{v}|,|\xi_{v}|,|\sigma_{v}|\ll 1$ and the end of the inflation corresponds to
$|x_{f}|\simeq \left(\frac{\tilde{C}^{2}_{0}}{8M^{2}\tilde{C}^{2}_{4}}\right)^{\frac{1}{6}}$ where $x_{f}=\phi_{f}-\phi_{0}$.
In this context equation of state parameter can be expressed as
\begin{equation}\begin{array}{lll}\label{state}
\omega(x)=\frac{P(x)}{\rho(x)}
=\left[\frac{-\tilde{C}^{2}_{4}x^{8}+\frac{16}{3}\tilde{C}^{2}_{4}M^{2}x^{6}-2\tilde{C}_{0}
\tilde{C}_{4}x^{4}-\tilde{C}^{2}_{0}}{\tilde{C}^{2}_{4}x^{8}+\frac{16}{3}\tilde{C}^{2}_{4}M^{2}x^{6}+2\tilde{C}_{0}
\tilde{C}_{4}x^{4}+\tilde{C}^{2}_{0}}\right]
\end{array}\end{equation}
\begin{figure}[htb]
{\centerline{\includegraphics[width=7cm, height=5cm]{state.eps}}}
\caption{Variation of equation of state parameter($\omega(x)$) versus shifted inflaton field ($x$)} \label{figVr38}
\end{figure}
which implies the energy scale of MSSM inflation is $\tilde{C}^{\frac{1}{4}}_{0}\sim(0.409-1.301)\times 10^{-9}M$
explicitly shown in the allowed region in Fig(\ref{figVr38}).
The number of e-foldings are defined \cite{lyth} for our model in the limit $\tilde{C}_{0}\gg\tilde{C}_{4}$ as
\begin{equation}\begin{array}{ll}\label{uit}{\cal N}=\log\left(\frac{a(t_{f})}{a(t_{i})}\right)\simeq \frac{1}{M^{2}}\int^{x_{i}}_{x_{f}}\frac{V}{V^{'}}dx
\simeq\frac{\tilde{C}_{0}}{8M^{2}\tilde{C}_{4}}\left\{\frac{1}{x^{2}_{f}}-\frac{1}{x^{2}_{i}}\right\},\end{array}\end{equation}
Further, in this framework the expressions for amplitude of the scalar perturbation, tensor perturbation
and tensor to scalar ratio are given by
\begin{eqnarray}\label{x1}\Delta^{2}_{s}&=&
\frac{V^{3}}{75\pi^{2}M^{6}(V^{'})^{2}}|_{k=aH}
=-\frac{32\tilde{C}_{4}{\cal N}^{3}}{75\pi^{2}},
\\
\label{x2}\Delta^{2}_{t}&=&
\frac{V}{150\pi^{2}M^{4}}|_{k=aH}
=\frac{(\tilde{C}_{0}+\tilde{C}_{4}x^{4}_{\star})}{150\pi^{2}M^{4}},
\\
\label{x3} r &=&
16\left(\frac{\Delta^{2}_{t}}{\Delta^{2}_{s}}\right)
=-\frac{\tilde{C}_{0}}{4\tilde{C}_{4}M^{4}{\cal N}^{3}}.\end{eqnarray}
Here $x_{\star}$ represents the value of the inflaton field
at the horizon crossing. For our model expression for spectral index, running and running of the running reduces to the following form:
\begin{eqnarray}
\label{d1}
n_{s}-1
&=&\frac{3\tilde{C}_{0}}{32\tilde{C}_{4}M^{4}{\cal N}^{3}}-\frac{3}{{\cal N}}
,\\
\label{d11}n_{t}&=
\frac{\tilde{C}_{0}}{32\tilde{C}_{4}M^{4}{\cal N}^{3}},\\ \label{famm}
\alpha_{s
&=&\frac{3}{{\cal N}^{2}}-\frac{2}{3}(n_{s}-1)^{2},\\ \label{ramm}
\kappa_{s
&=&-\frac{6}{{\cal N}^{3}}-\frac{4}{{\cal N}^{2}}(n_{s}-1)+\frac{8}{9}(n_{s}-1)^{3}.\end{eqnarray}
\begin{figure}[htb]
{\centerline{\includegraphics[width=7cm, height=5cm]{actual.eps}}}
\caption{$~$ Variation of the$~
scalar power spectrum($\Delta_{s}$) vs scalar spectral index($n_{s}$)} \label{figVr8667}
\end{figure}
\begin{figure}[htb]
{\centerline{\includegraphics[width=7cm, height=5cm] {paraG.eps}}}
\caption{$~$ Parametric plot of the logarithmic scaled amplitude of the scalar fluctuation ($ln({\Delta}_{s})$)
vs logarithmic scaled amplitude
of the running of the spectral index ($\ln(|\alpha_{s}|)$)$~$
.} \label{figVr241}
\end{figure}
Fig(\ref{figVr8667}) depicts the behavior of the scalar power spectrum as a function of scalar spectral index.
For ${\cal N}=70$ the scalar spectral index is within the bounds of
{\it WMAP7+BAO+h0 data} for model {\it $\Lambda$CDM+sz+lens} \cite{wmap07}.
Fig(\ref{figVr241}) shows the
behavior of amplitude of scalar fluctuation as a function of
running of the spectral index. For the best fit value of $\tilde{C}_{0}=2.867\times 10^{-36}M^{4}$,
$\tilde{C}_{4}=-1.685\times 10^{-13}$ and ${\cal N}=70$ the cosmological parameters obtained from our model is
$\Delta^{2}_{s}=2.498\times 10^{-9}$, $\Delta^{2}_{t}=1.936\times 10^{-39}$, $n_{s}=0.957$, $n_{t}=-1.550\times 10^{-30}$,
$r=1.240\times 10^{-29}$, $\alpha_{s}=-0.612\times 10^{-3}$, $\kappa_{s}=1.749\times 10^{-5}$.
Further, we use the publicly available code CAMB \cite{camb} to verify our results directly with observation.
To operate CAMB at the pivot scale $k_0=0.002~{\rm Mpc}^{-1}$ the values of the initial parameter space
are taken for $\tilde{C}_{0}=2.867\times10^{-36}M^{4}$ and ${\cal N}=70$.
Additionally WMAP7 years dataset \cite{wmap07} for
$\Lambda$CDM background has been used in CAMB to obtain CMB angular power spectrum.
In Table\ref{tab3} we have given all the input parameters for CAMB.
Table\ref{tab4} shows the CAMB output, which is in good agreement with WMAP seven years data.
In Fig.\ref{figVr1181}
we have plotted CAMB output of CMB TT angular power spectrum
$C_l^{TT}$ for best fit with WMAP seven years data for scalar mode, which explicitly show
the agreement of our model with WMAP7 dataset.
\begin{table}[htb]
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}
\hline $H_0$ & $\tau_{Reion}$ &$\Omega_b h^2$& $\Omega_c h^2
$& $T_{CMB}$
\\
km/sec/MPc& & && K\\
\hline
71.0&0.09&0.0226&0.1120&2.725\\
\hline
\end{tabular}
\caption{Input parameters}\label{tab3}
\end{table}
\begin{table}[htb]
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}
\hline $t_0$ & $z_{Reion}$ &$\Omega_m$&$\Omega_{\Lambda}$&$\Omega_k$&$\eta_{Rec}$& $\eta_0$
\\
Gyr& & && &Mpc & Mpc\\
\hline
13.707&10.704&0.2670&0.7330&0.0&285.10&14345.1\\
\hline
\end{tabular}
\caption{Output obtained from CAMB}\label{tab4}
\end{table}
\begin{figure}[htb]
{\centerline{\includegraphics[width=9cm, height=6cm]{tt.eps}}}
\caption{Variation of CMB angular power spectrum$~$ $C^{TT}_{l}$ for
best fit and WMAP seven years data with the multipoles l for scalar mode} \label{figVr1181}
\end{figure}
Now in the context of any running mass model one can expand the
spectral index with the following parameterization \cite{kosowsky}:
\begin{equation} \label{ntwer}
n({\cal R}) = n_z (k_0) - \frac{\alpha_z(k_0)}{2!}\ln\left( k_0
{\cal R} \right) + \frac{\kappa_z(k_0)}{3!}\ln^2\left( k_0 {\cal R} \right)+.....
\end{equation}
with ${\cal R} \ll 1/k_0$, i.e. $\ln(k_0 {\cal R}) < 0$. This is identified to be the significant
contribution to the Primordial Black Hole (PBH) formation. Here the
parameterization index $ z:\left[s(scalar),t(tensor)\right]$ and the explicit form of
the first term in the above expansion is given by
\begin{equation}\label{runfunc} n_z (k_0)=
\left\{
\begin{array}{ll}
n_s (k_0)-1 & \mbox{if } z=s \\
n_t (k_0) & \mbox{if } z=t.
\end{array}
\right.
\end{equation}
Existence of the running and running of the running is
the key feature in the formation of PBH in the radiation dominated era just after inflation \cite{dress}
which could form CDM in the Universe. The
initial PBHs mass $\cal M_{\text{PBH}}$ is related to the particle horizon mass
${\cal M}$ by
$ {\cal M_{\text{PBH}}}={\cal M }\gamma= \frac{4\pi}{3}\gamma \rho {\cal H}^{-3}\,$
at horizon entry, ${\cal R}=(a{\cal H})^{-1}$. This is formed when the density
fluctuation exceeds the threshold for PBH formation given as in {\it Press--Schechter theory} by \cite{dress,Copeland}
\begin{equation} \label{fofm}
f( \geq{\cal M}) = 2 \gamma \int _{\varTheta_{\rm th}}^{\infty} {\cal\text{d}\varTheta\,}{\cal P(\varTheta; M(R))}
.
\end{equation}
Here $\cal P(\varTheta; M(R))$ is the {\it Gaussian probability distribution function} of the
linearized density field $\varTheta$ smoothed on a comoving scale $\cal R$ by
${\cal P(\varTheta;R)} = \frac {1} {\sqrt{2\pi} \varSigma_{\varTheta}({\cal R})} \exp\left( -\frac
{\varTheta^{2}} {2\varSigma_{\varTheta}^{2}({\cal R})} \right)$
where the standard deviation
\begin{equation} \label{sigma}
\varSigma_{\varTheta}({\cal R}) =\sqrt{ \int_{0}^{\infty}\dfrac{\text{d}k} {k} \exp\left(-k^{2}{\cal R}^{2} \right)\ {\Delta}^2_{\varTheta}(k)
}\,.
\end{equation}
For our model power spectrum for $\varTheta(k,\eta)$ is given by
\begin{equation}\begin{array}{ll}\label{psx1}{\Delta}^2_{\varTheta}(k,\eta)
=\frac{4J} {25}\left( \frac{k}{a{\cal H}}
\right)^4 {\Delta}^2_{s}(k,\eta)=\frac{8J\sqrt{3\tilde{C}_{0}}Gk^4\eta^4(1+k^2\eta^2)} {25M\pi^2\kappa_{s}(\eta)},
\end{array}\end{equation}
where $\kappa_{s}(\eta)=\frac{18\sqrt{3}M^3G^3}{\tilde{C}^{\frac{3}{2}}_{0}\left[\Phi(f)-MG\sqrt{\frac{3}{
\tilde{C}_{0}}}\ln\left(\frac{\eta a(t_f)\sqrt{\tilde{C}_{0}}}{\sqrt{3}M}\right)\right]^3}$ is the equivalent
expression for running of the running in terms of conformal time $\eta$ and $J=\frac{(1+w)^2} {(1+\frac{3}{5}w)^2}$.
Additionally we have used
$\Phi(f)=x^{-2}_{f}$ and $G=\frac{8\tilde{C}_{4}M}{\sqrt{3\tilde{C}_{0}}}$.
Substituting eqn(\ref{psx1}) in eqn(\ref{sigma}) and using eqn(\ref{ramm})
at the horizon
crossing we get $\varSigma_{\varTheta}({\cal R})=\sqrt{\frac{8\sqrt{3\tilde{C}_{0}}JG}{25M\pi^2\kappa_{s}}\Gamma[\frac{(n_S({\cal R})+3)}{2}]}$ for
$n_s({\cal R})>3$. Consequently eqn(\ref{fofm}) gives
$f( \geq{\cal M}) = \gamma\text{erf}\left[\frac{\varTheta_{\rm th}}
{\sqrt{2} \varSigma_{\varTheta}({\cal R})}\right]$. Here we fix $\gamma
\simeq 0.2$ during the radiation dominated era \cite{Carr2} for proper numerical estimations. In general
the mass of PBHs is expected to depend
on the amplitude, size and shape of the perturbations \cite{Niemeyer}.
As a consequence the PBH mass is given by \cite{dress}
${\cal M_{\text{PBH}}} = \gamma {\cal M_{\text{eq}}} (k_{\text{eq}}{\cal R})^2 \left( \frac
{g_{\ast,\text{eq}}} {g_{\ast}} \right)^{\frac{1}{3}}\, ,$
where the subscript ``eq'' refers to the matter--radiation
equality. Here we use
$g_{\star}=228.75$ (all degrees of freedom
in MSSM), while $g_{\ast,\text{eq}} = 3.36$
and $k_{eq} = 0.07 \Omega_{\text{m}} h^{2}\ \text{Mpc}^{-1}$ (Here we use $\Omega_{\text{m}}
h^{2} = 0.2670$ from the CAMB output). Consequently
the relation between comoving scale and the PBH mass in the context of MSSM is given by
\begin{equation} \label{R}
\frac {\cal R} {1\ \text{Mpc}} = 1.250 \times 10^{-23}\left(
\frac {\cal M_{\text{PBH}}} {1\ \text{g}} \right)^{\frac{1}{2}}\, .
\end{equation}
\begin{figure}[htb]
{\centerline{\includegraphics[width=7cm, height=5cm]{pbh.eps}}}
\caption{$~$ Variation of the$~
fraction of the energy density of the universe collapsing into PBHs as a function
of the PBH mass, for three different values of the threshold
$\varTheta_{th} = 0.3(dashed), 0.5(solid), 0.7(dotdashed)$.
} \label{figVr8167}
\end{figure}
Fig(\ref{figVr8167}) shows the behavior of {\it Press--Schechter mass function}
with respect to PBH mass. With the values of the parameters as obtained earlier, we have
${\cal M}_{PBH}\simeq 10^{13}gm$ and
the corresponding fractional energy density $f=0.170$.
Finally, the reheating temperature for our model turns out to be
$T_{rh}=\left(\frac{30x^{2}_{\star}M^{2}}{g_{\star}\pi^{2}}\right)^{\frac{1}{4}}
\sqrt[12]{1200\pi^{2}\tilde{C}^{2}_{4}\Delta^{2}_{s}}$.
For ${\cal N}=70$ it is estimated as $T_{rh}=2.114\times 10^{8}GeV$ which is
obviously significant input to choose the fine tuned initial conditions
for RGE flow discussed in the next section.
\section{One loop RG flow}
For the flat direction \textbf{QQQL,QuQd,QuLe,uude} the soft SUSY breaking masses can be expressed as
\begin{equation}\begin{array}{llll}\label{jkhjg}
(m^{2}_{\phi})_{\textbf{QQQL}}=\frac{1}{4}(m^{2}_{{\bf\tilde{Q}_{a}}}
+m^{2}_{{\bf\tilde{Q}_{b}}}+m^{2}_{{\bf\tilde{Q}_{c}}}+m^{2}_{{\bf\tilde{L}_{3}}}),
\\ (m^{2}_{\phi})_{\textbf{QuQd}}=\frac{1}{4}(m^{2}_{{\bf\tilde{Q}_{a}}}
+m^{2}_{{\bf\tilde{Q}_{b}}}+m^{2}_{{\bf\tilde{u}_{c}}}+m^{2}_{{\bf\tilde{d}_{3}}}),
\\ (m^{2}_{\phi})_{\textbf{QuLe}}=\frac{1}{4}(m^{2}_{{\bf\tilde{Q}_{a}}}
+m^{2}_{{\bf\tilde{u}_{b}}}+m^{2}_{{\bf\tilde{L}_{c}}}+m^{2}_{{\bf\tilde{e_{3}}}}),
\\ (m^{2}_{\phi})_{\textbf{uude}}=\frac{1}{4}(m^{2}_{{\bf\tilde{u}_{a}}}+m^{2}_{{\bf\tilde{u}_{b}}}
+m^{2}_{{\bf\tilde{d}_{c}}}+m^{2}_{{\bf\tilde{e}_{3}}}),
\end{array}
\end{equation}
where ${\bf1\leq a,b,c\leq 3}$ and ${\bf a\neq b\neq c}$. After neglecting the contribution from the all Yukawa couplings except from the top we can express the one-loop beta function as
\cite{nilles}
$\beta_{m^{2}_{a}}=\dot{\mu}m^{2}_{a}
=\frac{1}{8\pi^2}\left(m^2_{a}+|A^{33}_{U}|^2\right)\left(\lambda^{33}_{U}\right)^{2}-\frac{1}{2\pi^{2}}\sum^{3}_{\alpha=1}g^{2}_{\alpha}|\tilde{m}_{\alpha}|^{2}\textbf{X}_{\alpha a}$
where $\textbf{X}_{\alpha a}$ are the quadratic Casimir Group Invariants for the superfield $\Phi$, defined in terms
of Lie Algebra generators $T^{a}$ by
$(T^{\alpha}T^{\alpha})^{a}_{b}=\textbf{X}_{\alpha a}\delta^{a}_{b}$
and $\dot{\mu}=\mu\frac{\partial}{\partial\mu}$.
\begin{figure}[htb]
{\centerline{\includegraphics[width=7cm, height=5cm]{mass.eps}}}
\caption{Running of gaugino mass ($m_{i}(\mu)$)
in one loop RGE for MSSM with the logarithmic scale $\log_{10}\left(\mu\right)$. Here we have used
$\mu_{0}=2.6\times 10^{7} GeV$ , $m_{i}(\mu_{0})=7.546\times10^{-3}TeV$, $\zeta=1$ $\forall$ i.} \label{figVr18}
\end{figure}
\begin{figure}[htb]
{\centerline{\includegraphics[width=8cm, height=5.5cm]{mquqd.eps}}
\caption{Running of soft mass squared ratio $\left(\frac{m^{2}_{\phi}(\mu)}{m^{2}_{\phi}(\mu_{0})}\right)$ in one loop RGE for MSSM
with the logarithmic scale $\log_{10}\left(\mu\right)$ where
$\mu_{0}=2.6\times 10^{7} GeV$ for $\zeta=0.5,1,2$ for $n=4$ level $~${\bf QuQd} $\forall$ i. Similar plots can be obtained for
{\bf QuLe}, {\bf QQQL} and
{\bf uude} flat directions also $\forall$ i.} \label{figVr1568}
\end{figure}
\begin{figure*}
{\centerline{\includegraphics[width=6.4cm, height=5cm]{au.eps} \includegraphics[width=6.4cm, height=5cm]{ad.eps} \includegraphics[width=6.4cm, height=5cm]{ae.eps}}}
\caption{Running of trilinear A -term ratio $\left(\frac{A_{\beta}(\mu)}{A_{\beta}(\mu_{0})}\right)$ in one loop RGE for MSSM with the logarithmic scale $\log_{10}\left(\mu\right)$ where
$\mu_{0}=2.6\times 10^{7} GeV$, $\zeta=0.5({\bf dotdashed}),
1({\bf solid}),2({\bf dashed})$ and $\beta=1(U),2(D),3(E)$ for $n=4$ level $\forall$ i.} \label{figVr118}
\end{figure*}
In the context of MSSM \\ $\textbf{X}_{1a}=\frac{3{\bf Y}^{2}_{a}}{5}$(for each ${\bf\Phi}_{a}$ with weak hyper charge ${\bf Y}_{a}$),
\\$\textbf{X}_{2a}=\frac{3}{4}$ (for ${\bf\Phi}_{a}={\bf Q,L,H_{u},H_{d}}$),
\\ $~~~~~~= 0 $ (for ${\bf\Phi}_{a}={\bf \bar{u},\bar{d},\bar{e}}$),
\\$\textbf{X}_{3a}=\frac{4}{3}$ (for ${\bf\Phi}_{a}={\bf Q,\bar{u},\bar{d}}$),
\\ $~~~~~~= 0 $ (for ${\bf\Phi}_{a}={\bf L,\bar{e},H_{u},H_{d}}$),
\\where $\textbf{X}_{1a}$, $\textbf{X}_{2a}$ and $\textbf{X}_{3a}$ are applicable for ${\bf U(1)_{Y}}$,${\bf SU(2)_{L}}$ and
${\bf SU(3)_{C}}$ respectively.
So for the flat direction content \textbf{QQQL,QuQd,QuLe,uude} we have the following beta functions:\\ \newpage
(a) \textbf{For Soft mass}:\\
$\dot{\mu}(m^{2}_{\phi})_{\textbf{QQQL}}= \frac{1}{8\pi^2}\left(3m^{2}_{Q_{3}}+m^{2}_{U_3}+|A^{33}_{U}|^2\right)\left(\lambda^{33}_{U}\right)^{2}$
\\$~~~~~~~~~~~~~~~~~~~-\frac{1}{8\pi^{2}}\left(3g^{2}_{2}|\tilde{m}_{2}|^{2}+4g^{2}_{3}|\tilde{m}_{3}|^{2}\right),$
\\
$\dot{\mu}(m^{2}_{\phi})_{\textbf{QuQd}}=\frac{1}{2\pi^2}\left(m^{2}_{Q_{3}}+m^{2}_{U_3}\right)\left(\lambda^{33}_{U}\right)^{2}$\\
$~~~~~~~~~~~~~~~~~~~~~~-\frac{1}{8\pi^{2}}\left(\frac{3}{2}g^{2}_{2}|\tilde{m}_{2}|^{2}+\frac{16}{3}g^{2}_{3}|\tilde{m}_{3}|^{2}\right),$
\\
$\dot{\mu}(m^{2}_{\phi})_{\textbf{QuLe}}=\frac{3}{8\pi^2}\left(m^{2}_{Q_3}+m^{2}_{U_3}\right)\left(\lambda^{33}_{U}\right)^{2}$\\
$~~~~~~~~~~~~~~~~~~~~~~-\frac{1}{8\pi^{2}}\left(\frac{3}{2}g^{2}_{2}|\tilde{m}_{2}|^{2}+\frac{8}{3}g^{2}_{3}|\tilde{m}_{3}|^{2}\right),$
\\
$\dot{\mu}(m^{2}_{\phi})_{\textbf{uude}}=\frac{1}{2\pi^2}\left(m^{2}_{Q_{3}}+m^{2}_{U_3}\right)\left(\lambda^{33}_{U}\right)^{2}
-\frac{1}{2\pi^{2}}g^{2}_{3}|\tilde{m}_{3}|^{2},$\\
(b) \textbf{For Trilinear A- term}:\\
$~~~~~~\dot{\mu}A^{aa}_{D}=\delta_{b3}\left(\lambda^{33}_{U}\right)^{2}\frac{A^{33}_{U}}{8\pi^{2}}\\
~~~~~~~~~~~~~~~-\frac{1}{4\pi^{2}}\left(\frac{7}{18}g^{2}_{1}|\tilde{m}_{1}|^{2}+
\frac{3}{2}g^{2}_{2}|\tilde{m}_{2}|^{2}+\frac{8}{3}g^{2}_{3}|\tilde{m}_{3}|^{2}\right),$
\\
$~~~~~~\dot{\mu}A^{ab}_{U}=\frac{3(1+\delta_{a3})A^{33}_{U}}{8\pi^{2}}\left(\lambda^{33}_{U}\right)^{2}\\
~~~~~~~~~~~~~~~-\frac{1}{4\pi^{2}}\left(\frac{13}{18}g^{2}_{1}|\tilde{m}_{1}|^{2}+\frac{3}{2}g^{2}_{2}|
\tilde{m}_{2}|^{2}+\frac{8}{3}g^{2}_{3}|\tilde{m}_{3}|^{2}\right),$
\\
$~~~~~~\dot{\mu}A^{aa}_{E}=
-\frac{1}{4\pi^{2}}\left(\frac{3}{2}g^{2}_{1}|\tilde{m}_{1}|^{2}+\frac{3}{2}g^{2}_{2}|\tilde{m}_{2}|^{2}\right),$\\
(c) \textbf{For Fourth level Yukawa coupling}:\\
$~~~~~~\dot{\mu}\lambda^{aa}_{U}=\frac{3(1+\delta_{a3})}{8\pi^{2}}\left(\lambda^{33}_{U}\right)^{3}
-\frac{\lambda^{aa}_{U}}{4\pi^{2}}\left(\frac{13}{18}g^{2}_{1}+\frac{3}{2}g^{2}_{2}+\frac{8}{3}g^{2}_{3}\right),$
\\
$~~~~~~\dot{\mu}\lambda^{ab}_{D}=\delta_{b3}\left(\lambda^{33}_{U}\right)^{2}\frac{\lambda^{ab}_{D}}{8\pi^{2}}
-\frac{\lambda^{ab}_{D}}{4\pi^{2}}\left(\frac{7}{18}g^{2}_{1}+\frac{3}{2}g^{2}_{2}+\frac{8}{3}g^{2}_{3}\right),$
\\
$~~~~~~\dot{\mu}\lambda^{aa}_{E}=
-\frac{\lambda^{aa}_{E}}{4\pi^{2}}\left(\frac{3}{2}g^{2}_{2}+\frac{3}{2}g^{2}_{3}\right)$\\
where all the superscript a and b represent generation or family indices run from 1 to 3 physically representing the
first, second and third generation respectively. For the one-loop renormalization of gauge couplings and gaugino masses, one has in general
\begin{equation}\begin{array}{llll}\label{poi}\beta_{g_{\alpha}}=\dot{\mu}g_{\alpha}=\frac{g^{3}_{\alpha}}{16\pi^{2}}\left[\Sigma_{a}\textbf{I}_{\alpha a}-3\textbf{X}_{\alpha G}\right],\\
\beta_{m_{\alpha}}=\dot{\mu}m_{\alpha}=\frac{g^{2}_{\alpha}m_{\alpha}}{8\pi^{2}}\left[\Sigma_{a}\textbf{I}_{\alpha a}-3\textbf{X}_{\alpha G}\right]
\end{array}
\end{equation}
where $\textbf{X}_{\alpha G}$ quadratic Casimir invariant of the group $[$ 0 for $\bf{U(1)}$ and $\bf{N}$ for $\bf{SU(N)}$ $]$, $\textbf{I}_{\alpha a}$ is the Dynkin
index of the chiral supermultiplet $\bf{\Phi}_{a}$ $[$ normalized to $\frac{1}{2}$ for each fundamental representation of $\bf{SU(N)}$ and to
$3\bf{Y}^{2}_{a}/5$ for $\bf{U(1)}_{Y}$ $]$. For the above mentioned flat direction
the running of gauge couplings ($g_{i}(\mu)$) and gaugino masses ($m_{i}(\mu)$) obey \cite{nilles},
$\dot{\mu}g_{i}=
\frac{d_{i}}{2}g^{3}_{i},~~~~~~
\dot{\mu}\left(\frac{m_{i}}{g^{2}_{i}}\right)=
0~\forall~~ i$
where for $i=1({\bf U(1)_{Y}}),2({\bf SU(2)_{L}}),3({\bf SU(3)_{C}})$ here $d_{1}=\frac{11}{8\pi^{2}}$,$d_{2}=\frac{1}{8\pi^{2}}$,$d_{3}=-\frac{3}{8\pi^{2}}$
which is the simpler version of the equation(\ref{poi}).
Now to show explicitly that the contributions from the top Yukawa coupling ($\lambda^{33}_{U}$) are very small
for an induced electroweak group $\textbf{G}_{EW}$=$\textbf{SU(2)}_{L}$ $\otimes$ $\textbf{U(1)}_{Y}$ breakdown,
let us start with the Higgs potential \cite{nilles,martin}
\begin{equation}
\begin{array}{lllll}\label{yuki}V_{ Higgs}({\bf H,\bar{H}})=m^2_{1}|{\bf H}|^2+m^2_{2}|{\bf\bar{H}}|^2+m^{2}_{3}
\left({\bf H{\bar{H}}}+{\bf H^{\dagger}\bar{H^{\dagger}}}\right)\\
~~~~~~~~~~~~~~~~~~~~~~~~~~~~~+\frac{1}{8}\left(g^2_{1}+g^2_{2}\right)\left[|{\bf H}|^2-|{\bf\bar{ H}}|^2\right]^2,\end{array}
\end{equation}
where ${\bf H}=H_{u}$ and ${\bf \bar{H}}=H_{d}$ represent the Higgs superfields and
the relative vev of the two
Higgses are given by
\begin{equation}\label{desy}
\begin{array}{lllllll}
v=\sqrt{{\langle {\bf H}\rangle}^2+{\langle {\bf {\bar H}}\rangle}^2}
\\~~\displaystyle=\sqrt{\frac{2\left[m^2_{1}-m^2_{2}-\left(m^2_{1}+m^2_{2}\right)cos(2{\bf\theta})\right]}{\left(g^2_{1}+g^2_{2}\right)cos(2{\bf\theta})}}
\end{array}
\end{equation}
with $tan(\theta)=\frac{{\langle {\bf {\bar H}}\rangle}}{{\langle {\bf H}\rangle}}$. Here $\theta$ represents
an angular parameter which parameterizes MSSM. For the sake of convenience
let us now write $cos(2\theta)$ appearing in equation(\ref{desy}) introducing new parameterization as \cite{iban}
$cos(2\theta)=\frac{w^2-1}{w^2+1}$ where $w=\frac{\frac{\langle {\bf H}\rangle}{v}}{\sqrt{1-\left(\frac{\langle {\bf H}\rangle}{v}\right)^2}}$.
Consequently the top Yukawa coupling can be expressed as $\lambda^{33}_{U}=\frac{m_{U}}{v sin(\theta)}$ where $0\leq\theta<\frac{\pi}{2}$
and the top mass 43 GeV$\leq m_{U}\leq 170$ GeV$\ll\mu_{GUT}$
comes from the RG flow \cite{nilles}. It is evident from the above parameterization \cite{iban,kaku,nano} that
as $w\rightarrow 1$, $\theta\rightarrow \frac{\pi}{4}$ which implies ${\langle {\bf H}\rangle}$ and ${\langle {\bf\bar{ H}}\rangle}$ is very large and have the same order of magnitude.
As a result the relative vev $v$ is also large and the top Yukawa coupling is very very small
for which one can easily neglect it from the RG flow at the energy scale of MSSM inflation as mentioned earlier.
The consequence of the large vev of Higgs field can be taken care of by introducing strongly interacting gauge group ${\bf G_{NEW}= G_{S}\otimes SU(3)_{C}}$
and its superconformal version ${\bf G_{SCONF}=SU(3)_{SC}\otimes SU(3)_{C}}$ \cite{kobayashi}.
In table(\ref{tab8}) we have tabulated the numerical values of vev of ${\bf H}$ and ${\bf \bar{H}}$, the angular parameter $\theta$,
tan($\theta$), $w$, the top mass $m_{U}$ and
the top Yukawa coupling $\lambda^{33}_{U}$ contributing to the parameter space of MSSM for the $n=4$ level flat directions
\textbf{QQQL,QuQd,QuLe} and \textbf{uude}. It should be noted that appearance of large vev of Higgses as mentioned in table(\ref{tab8}) can easily be interpreted
when Einstein Hilbert term appears in the total action of the theory at lowest order approximation \cite{kalo} which
is our present consideration. Consequently the contributions from the hard cutoff is sub-leading due to the soft conformal symmetry
breaking. This leads to small top Yukawa coupling in the restricted parametric space
of MSSM characterized by the phenomenological bound: 43 $GeV\leq m_{U} \leq$ 170 $GeV$, $1.006\leq tan(\beta)\leq 1.025$
for the n=4 flat directions.
Neglecting all the sub-leading contributions arising from the top Yukawa coupling in the restricted parameter space of the MSSM, the solutions of these RGE for n=4 level flat directions can be written as
\begin{equation}\begin{array}{llll}\label{s1}
g_{i}(\mu)=
\frac{g_{i}(\mu_{0})}{\sqrt{1-d_{i}g^{2}_{i}(\mu_{0})\ln\left(\frac{\mu}{\mu_{0}}\right)}},
\\
m_{i}(\mu)=m_{i}(\mu_{0})\left(\frac{g_{i}(\mu)}{g_{i}(\mu_{0})}\right)^{2},
\\
\Delta m^{2}_{\phi}=
\sum^{3}_{i=1}f_{F}^{i}\Delta m^{2}_{i},
\\
\Delta A^{ab}_{\beta}=
\frac{1}{2}\sum^{3}_{i=1}(C_{\beta}^{i})^{ab}\Delta m_{i},
\\
\lambda^{ab}_{\beta}(\mu)=
\lambda^{ab}_{\beta}(\mu_{0})\prod^{3}_{i=1}\left(\frac{g_{i}(\mu_{0})}{g_{i}(\mu)}\right)^{(C_{\beta}^{i})^{ab}},\end{array}\end{equation}
Here $g_{i}(\mu_{0})$, $m_{i}(\mu_{0})$, $A_{\beta}(\mu_{0})$, $m_{\phi}(\mu_{0})$ and $\lambda_{\beta}(\mu_{0})$ represent the value of the gauge couplings, gaugino masses,
trilinear couplings, soft SUSY braking masses and Yukawa couplings at the characteristic scale $\mu_{0}$. In equation(\ref{s1}) we
have used the following shorthand notations:
$~~~~~~~~~~~~~\Delta A_{\beta}=A_{\beta}(\mu)-A_{\beta}(\mu_{0})$,\\ $~~~~~~~~~~~~~~~~~\Delta m_{i}=m_{i}(\mu_{0})-m_{i}(\mu)$,\\
$~~~~~~~~~~~~~~~~~\Delta m^{2}_{\phi}=m^{2}_{\phi}(\mu)-m^{2}_{\phi}(\mu_{0})$,\\ $~~~~~~~~~~~~~~~~~\Delta m^{2}_{i}=m^{2}_{i}(\mu_{0})-m^{2}_{i}(\mu)$,\\
where the $\beta$ indices 1,2,3 represent U, D, E
respectively.
\begin{table}[htb]
\begin{tabular}{|c|c|c|c|c|}
\hline ${\bf f_{F}^{i}}$ & ${\bf i=1(U(1)_{Y})}$ &${\bf i=2(SU(2)_{L})}$&${\bf i=3(SU(3)_{C})}$\\
\hline
F=1({\bf QQQL})&0&$\frac{3}{2}$&-$\frac{2}{3}$\\
\hline
F=2({\bf QuQd})&0&$\frac{3}{4}$&-$\frac{8}{9}$\\
\hline
F=3({\bf QuLe})&0&$\frac{3}{4}$&-$\frac{4}{9}$\\
\hline
F=4({\bf uude})&0&0&-$\frac{2}{3}$\\
\hline
\end{tabular}
\caption{Entries of $f_{F}^{i}$ matrix obtained from the solution of RGE}\label{tab5}
\end{table}
\begin{table}[htb]
\begin{tabular}{|c|c|c|c|c|}
\hline ${\bf (C_{\beta}^{i})^{ab}}$ & ${\bf i=1(U(1)_{Y})}$ &${\bf i=2(SU(2)_{L})}$&${\bf i=3(SU(3)_{C})}$\\
\hline
$\beta$=1({\bf U}),a=b&$\frac{26}{99}$&6&-$\frac{32}{9}$\\
\hline
$\beta$=2({\bf D}),a=b&$\frac{14}{99}$&6&-$\frac{32}{9}$\\
\hline
$\beta$=3({\bf E}),$a\neq b$&$\frac{6}{11}$&6&0\\
\hline
\end{tabular}
\caption{Entries of $(C_{\beta}^{i})^{ab}$ matrix obtained from the solution of RGE}\label{tab6}
\end{table}
In equation (\ref{s1}) $f_{F}^{i}$ and $(C_{\beta}^{i})^{ab}$ are $(4\times 3)$ and $(3\times 3)$ matrices whose entries
are tabulated in Table(\ref{tab5}) and Table(\ref{tab6}) respectively. It is obvious from the RGE that $\beta=1,2$ implies $a=b$ and $\beta=3$ implies $a\neq b$.
\begin{figure}[htb]
{\centerline{\includegraphics[width=7.5cm, height=5cm]{lamb.eps}}}
\caption{Running of the ratio of the Yukawa coupling $\left(\frac{\lambda_{\beta}(\mu)}{\lambda_{\beta}(\mu_{0})}\right)$
in one loop RGE for MSSM with the logarithmic scale $\log_{10}\left(\mu\right)$. Here we have used
$\mu_{0}=2.6\times 10^{7} GeV$ and $\beta=1(U),2(D),3(E)$ $\forall$ i.} \label{figVr178}
\end{figure}
Using the solutions of RGE along with the approximation that the running of the gaugino masses and gauge couplings
is very very small we get:
\begin{equation}\begin{array}{lll}\label{cg1} D_{1}=
-\frac{1}{8\pi^{2}}\sum^{3}_{i=1}J_{i}\left(\frac{m_{i}}{m_{\phi_{0}}}\right)^{2}g^{2}_{i}(\mu_{0}),
\\
D^{\beta}_{2}=
-\frac{1}{4\pi^{2}}\sum^{3}_{i=1}K^{\beta i}\left(\frac{m_{i}}{A_{0}}\right)g^{2}_{i}(\mu_{0}),\end{array}\end{equation}
where we have $J_{1}=0$,$J_{2}=3$ and $J_{3}=4$ for $i=1, 2, 3$ and all the entries of $K^{\beta i}$ $(3\times 3)$ matrix are tabulated in
table(\ref{tab7}).
\begin{table}[htb]
\begin{tabular}{|c|c|c|c|c|}
\hline ${\bf K^{\beta i}}$ & ${\bf i=1(U(1)_{Y})}$ &${\bf i=2(SU(2)_{L})}$&${\bf i=3(SU(3)_{C})}$\\
\hline
$\beta$=1({\bf U})&$\frac{13}{18}$&$\frac{3}{2}$&$\frac{8}{3}$\\
\hline
$\beta$=2({\bf D})&$\frac{7}{18}$&$\frac{3}{2}$&$\frac{8}{3}$\\
\hline
$\beta$=3({\bf E})&$\frac{3}{2}$&$\frac{3}{2}$&0\\
\hline
\end{tabular}
\caption{Entries of $K^{\beta i}$ matrix}\label{tab7}
\end{table}
In this context the subscript `0' represents the values of parameters at the high scale $\mu_{0}$. As discussed in section III, constraining only $D_{1}$ and $D^{\beta}_{2}$ is sufficient here. Eqn(\ref{con2}) provides an
extra constraint relation which restricts the parameters further leading to more precise information in RG flow.
For universal boundary conditions, the high scale is identified to be
the GUT scale $\mu_{GUT} \approx 3 \times
10^{16}$~GeV, ${\tilde m_{1}}(\mu_{GUT}) = {\tilde m_{2}}(\mu_{GUT}) ={\tilde m_{3}}(\mu_{GUT})= {\tilde m}$,
$A_{E}(\mu_{GUT})=A_{U}(\mu_{GUT})=A_{D}(\mu_{GUT})=A_{0}$
and $g_{1}\approx 0.56$, $g_{2}\approx
0.72$, $g_{3}\approx 0.85$. Now depending upon the different phenomenological situations the $n=4$ level flat directions
are divided into two classes. The first class deals with ${\bf QuQd, QuLe}$
which is lifted completely at $n=4$ level. The other class which is lifted by higher dimensional operators
deals with ${\bf uude, QQQL}$. Most importantly ${\bf uude, QQQL}$ take part in the proton
decay (${\it p\rightarrow\pi^{0}e^{+}}$, ${\it p\rightarrow\pi^{+}\nu_{e}}$ etc.) \cite{pdec} which introduces a stringent constraint
on the Yukawa coupling $\lambda_{0}$ at $n=4$ level. Additionally the neutrino-antineutrino oscillation data
restricts $\lambda_{0}$ again.
Then we just use RG equations along with these restrictions to run the coupling constants and masses to the scales as mentioned in
table(\ref{tab8}) with $M=2.4\times10^{18}$~GeV.
\begin{center}
\begin{table*}
{\small
\hfill{}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|c|c|}
\hline ${\bf Flat~}$ & ${\bf\mu_{0}=\phi_{0}}$&$A_{0,tree} $&${\bf m_{\phi_{0}}}$&${\bf\langle {\bf H}\rangle }$& ${\bf\langle {\bf \bar{H}}\rangle }$&$\theta$&$tan(\theta)$&$w$&$v$&$m_{U}$ & ${\bf \lambda^{33}_{U}=\lambda_{0}}$\\
${\bf direction}$ & ${ GeV}$&${ GeV}$ &${ GeV}$&${ GeV}$&${ GeV}$&in $\deg$& & &${ GeV}$&${ GeV}$&${ GeV}$\\
\hline
${\bf QuLe}$&$2.6 \times
10^{7} $&$36.967$&$7.546$&$0.200\times 10^{16}$ & $0.458\times 10^{16}$& 45.171&1.006 &0.994 &$0.500\times 10^{16} $& 43&$1.212\times 10^{-14} $\\
\hline
${\bf QuQd}$&$2.6 \times
10^{7} $&$36.967$&$7.546$&$ 0.450\times 10^{16} $ & $0.423\times 10^{16}$ & 45.370 &1.013 & 0.987 &$0.601\times 10^{16}$ &170 &$7.106\times 10^{-14} $\\
\hline
${\bf QQQL}$&$1.344\times
10^{14}$&$892\times 10^{3}$&$182\times 10^{3}$&$0.188\times 10^{8}$&$0.124\times 10^{8}$&45.707 & 1.025 &0.975 &$0.226\times 10^{8}$& 80 &$4.945\times 10^{-6}$\\
\hline
${\bf uude}$&$2.896\times
10^{13}$&$4.142\times10^{6}$&$845\times 10^{3}$&$0.174\times 10^{6}$ &$0.157\times 10^{6}$ & 45.549 & 1.019 &0.981 &$0.235\times 10^{6}$ &135 & $8.047\times 10^{-4}$\\
\hline
\end{tabular}}
\hfill{}
\caption{MSSM parameter values obtained from RG flow for n=4 level flat directions}\label{tab8}
\end{table*}
\end{center}
Considering all these values we obtain effectively\\
$~~~~~~~~~~~~~~~~~~~~~D_{1} \approx -0.056 \zeta^2, $\\
$~~~~~~~~~~~~~~~~~~~~~D^{1}_{2} \approx -0.074 \zeta,$\\
$~~~~~~~~~~~~~~~~~~~~~D^{2}_{2} \approx -0.071 \zeta,$\\
$~~~~~~~~~~~~~~~~~~~~~D^{3}_{2} \approx -0.031 \zeta,$\\
$~~~~~~~~~~~~~~~~~~~~~D^{1}_{3}= D^{2}_{3}=D^{3}_{3}\approx -0.048-0.168\zeta^{2}$,\\
where $\zeta = m/ m_{\phi}$ is calculated at the GUT scale. Typically the running
based on gaugino loops alone results in negative values of
$D_{i}\forall i$. Positive values can be obtained when one includes the Yukawa couplings,
practically the top Yukawa, but the order of magnitude remains the same.
The choice of fine tuned initial conditions directly shows
more fine tuning is required compared to other models. It is a straightforward exercise to verify that
even if one considers all the flat directions at $n=4$ level one will arrive at the potential eqn.(\ref{hgkl})
with same $\tilde{C}_{0}$ and $\tilde{C}_{4}$. This is precisely what we have done in this paper.
The results of RG flow have been demonstrated in figs(\ref{figVr18})-(\ref{figVr178}).
In fig(\ref{figVr18}) and fig(\ref{figVr1568})
`{\it dashed}', `{\it solid}' and `{\it dotdashed}' line represents ${\bf U(1)_{Y}}$, ${\bf SU(2)_{L}}$ and ${\bf SU(3)_{C}}$
gauge group content respectively. Fig(\ref{figVr18})-fig(\ref{figVr178}) explicitly showing
the behavior of the RGE flow of gaugino masses, soft SUSY
breaking mass, trilinear couplings and Yukawa couplings respectively.
Additionally fig(\ref{figVr18})-fig(\ref{figVr118}) give consistent GUT scale unification.
\section{Summary and outlook}
In this article we have proposed a model of inflation
in the framework of MSSM with new flat directions using saddle point mechanism.
We have demonstrated how we can construct the
effective inflationary potential in the vicinity of the {\it saddle point} starting
from $n=4$ level superpotential for the flat direction content
\textbf{QQQL,QuQd,QuLe} and \textbf{uude} for MSSM. The effective inflaton potential
around saddle point, resulting from the non-vanishing
fourth derivative of the original potential, has then been utilized in estimating for
the observable parameters and confronting them with
WMAP7 dataset using the publicly available code CAMB, which reveals consistency of our model with latest observations.
We have then explored the possibility of Primordial Black Hole formation from the
running-mass model by estimating the mass of PBH.
Subsequently, we have engaged ourselves in finding out the effective parameter space
and the constants appearing in the {\it saddle point} analysis for the MSSM inflation
by solving the one loop RGE. It is worth mentioning that the RGE flow of fourth
level MSSM is exactly solvable in this context and we hope that
all the numerics can be tested in the LHC or any linear
collider in near future.
Consequently we conclude
that fourth level MSSM inflation confronts extremely well
with WMAP7 within a certain parameter space obtained from one loop MSSM RGE flow.
A detailed survey of
RG flow with two loop beta function,
inflection point inflation \cite{anupam3} for $n=4$ level
MSSM candidates, sensitivity
in the neighborhood of the saddle point with the one loop corrected
potential, the effect of quantum Coleman De Luccia tunneling \cite{Coleman} and
the inflationary model building of MSSM derived from string theory via braneworld using several compactification schemes remain an
open issue, which may even provide interesting signatures of MSSM
inflation. We hope to address some of these issues in due course.
\section*{Acknowledgments}
SC thanks H. P. Nilles, B. K. Pal and I. Singh
for discussions and Council of Scientific and
Industrial Research, India for financial support through Junior
Research Fellowship (Grant No. 09/093(0132)/2010). SP is partially supported
by the Alexander von Humboldt Foundation Germany, the SFB-Tansregio TR33 ``The Dark Universe''
(Deutsche Forschungsgemeinschaft) and the European Union 7th
network program ``Unification in the LHC era''
(PITN-GA-2009-237920).
We also acknowledge illuminating discussions with A. Mazumdar which helped in improvement of the article.
|
\section{Introduction}
At imaginary chemical potential QCD has an interesting symmetry, known as the Roberge-Weiss (RW) symmetry~\cite{RW}: As a remnant of the \ensuremath{\mathbb Z_3}{} symmetry of the pure SU(3) gauge theory, certain shifts in the imaginary chemical potential $\mu = i \theta T$ can be undone by a \Z\ transformation{} leading to a periodicity of $\theta \rightarrow \theta+ 2\pi k/3$ with integer $k$ in thermodynamic quantities like the pressure.
At large temperatures the system undergoes a first-order transition jumping between different \ensuremath{\mathbb Z_3}{} sectors when crossing $\theta = (2 k+1) \,\pi/3$ for fixed temperature. Due to the periodicity $\theta$-even quantities show a cusp, whereas $\theta$-odd quantities have a jump. For low temperatures this transition is a crossover. In between there must be an endpoint of the RW transition which can be of first or second order. If the transition along $\theta = \pi/3$ ends in a first-order transition, there must be first-order lines departing from it implying that the endpoint is a triple point. As first-order phase transitions and second-order endpoints might influence the phase structure at real chemical potential, this warrants further studies.
Recent lattice QCD simulations at imaginary chemical potential for two and three quark flavors have shown that the order of the RW endpoint depends on the quark masses~\cite{DEliaRWEndpoint, PhilipsenRWEndpoint}. For low and high masses, the transition is of first order. A first-order transition at large quark masses is to be expected from the limit of pure SU(3) gauge theory. In the intermediate mass range the transition changes to second order with tricritical points in between.
Since lattice studies are hampered by the sign problem and are computationally very demanding, it is worth studying these aspects in effective models. In the Polyakov-loop extended Nambu--Jona-Lasinio (PNJL) model, which can be applied for real as well as at imaginary chemical potentials, we thus investigate the phase structure in the $\mu^2-T$-plane. At imaginary quark chemical potential the PNJL model also features the RW symmetry and we find the RW periodicity as well as the RW phase transition.
The PNJL model at imaginary chemical potential has already been investigated by Sakai et.\,al.~\cite{Sakai0902,Sakai0904}.
In a two-flavor PNJL model we extend their work and study the order of the RW phase transition endpoint for different Polyakov-loop potentials and analyze its dependence on the relative strength of the potentials~\cite{MScThesis}. This is done in two ways: Since quarks with larger mass have a smaller contribution to the pressure, increasing the quark mass makes the gluonic part more important. Alternatively, we directly change the prefactor of the gluonic contribution.
\section{Model}
We employ the PNJL model for two light quark flavors at real and imaginary chemical potential in mean-field approximation following the standard procedures. The Lagragian is given by
\begin{align*}
\mathcal{L}_\text{PNJL} = &\bar\psi \left( i \gamma_\mu D^\mu - m_0 \right)\psi
+\frac{g_S}{2}\left[ (\bar\psi \psi)^2 + (\bar\psi i \gamma_5 \tau_a \psi)^2 \right]+ \mathcal{U}(\Phi,\bar\Phi)
\end{align*}
with quark fields $\psi$, covariant derivative $D^\mu$, bare quark mass $m_0$, coupling constant $g_S$ of the four-quark interaction and the Polyakov-loop potential $\mathcal{U}$, modelling the gluonic contributions which depends on the Polyakov-loop variables $\Phi$ and $\bar\Phi$. Parameters are taken from~\cite{Sakai0902}.
The \textit{extended \Z\ transformation}~\cite{Sakai0904} is given by
\begin{gather*}
\begin{split}
\theta&\quad\rightarrow\quad \theta+ 2\pi k/3 \\
\Phi &\quad\rightarrow\quad \Phi \exp{[-i 2\pi k/3]} \quad\text{with}\quad k \in \mathbb{Z}\text.
\end{split}
\label{eq:eZt}
\end{gather*}
A convenient definition is the \textit{modified Polyakov loop}, $\Psi = \Phi \exp{[i\theta]}$, which is then invariant under the extended \Z\ transformation. It can easily be shown that the PNJL model is invariant under the extended \Z\ transformation{} and thus possesses the RW periodicity at imaginary chemical potential.
\section{Results}
We start with a logarithmic form of the Polyakov-loop potential~\cite{RRW06},
\begin{equation*}
\frac{\mathcal{U}_\text{log}}{T^4} = -\frac{a(T)}{2} \Phi\bar\Phi + b(T) \log\left[ 1-6\Phi\bar\Phi + 4(\Phi^3+{\bar\Phi}^3) - 3(\Phi\bar\Phi)^2 \right] \text.
\end{equation*}
We show the behavior of the order parameters at fixed $\theta=0$ and $\theta=\pi/3$ in Fig.~\ref{fig:orderparameters01}. Along $\theta=\pi/3$ which is in the middle of the period we find a jump in the absolute value and the phase of the Polyakov loop.
The dependence on $\theta$ at fixed temperatures close to the RW transition is displayed in Fig.~\ref{fig:orderparametersRW}. At temperatures higher than the transition temperature $T_{RW}$ the phase has a jump and the absolute value a cusp when crossing the RW phase transition, signalling the jump from one \ensuremath{\mathbb Z_3}{} sector to another. At temperatures slightly lower than $T_{RW}$ we however find two jumps in the phase and also in the absolute value. In addition the chiral condensate picks up the same discontinuities as the absolute value of the Polyakov loop. For even lower temperatures all transitions are continuous.
We summarize these findings in the PNJL phase diagram shown in Fig.~\ref{fig:pd}. Crossover lines are determined by the inflection point of the Polyakov-loop absolute value as a function of temperature. Chiral crossover lines are ommitted as they are not relevant for our current analysis.
\begin{figure}
\centering
\includegraphics[width=0.4\textwidth]{mu_0.eps}\hspace{1cm}
\includegraphics[width=0.4\textwidth]{theta_1.eps}
\caption{Modified Polyakov-loop variables and the normalized chiral condensate $\sigma/\sigma_0$ at $\theta= 0$ (left) and $\theta= \pi/3$ (right) as functions of temperature. The phase of $\Psi$ vanishes at $\theta=0$ and only the positive branch is shown at $\theta=\pi/3$.}
\label{fig:orderparameters01}
\end{figure}
\begin{figure}
\includegraphics[width=\textwidth]{legsorderparams.eps}
\caption{Dependence of the modified Polyakov-loop variables and the normalized chiral condensate on $\theta$ for different temperatures around $T_{RW}=190.3$ MeV (red solid: $T=185$~MeV, green dashed: $188$~MeV, blue dotted: $191$~MeV).}
\label{fig:orderparametersRW}
\end{figure}
\begin{figure}
\includegraphics[width=0.49\linewidth]{pd.eps}
\includegraphics[width=0.50\linewidth]{pd2.eps}
\caption{Phase diagram in the $\theta-T$ (left) plane and the $\mu^2-T$ (right) plane. Red solid lines denote first-order RW/deconfinement transitions, green dashed lines show the deconfinement crossover, and the blue solid line at real chemical potential denotes the chiral first-order transition. The diamonds represent second-order endpoints.}
\label{fig:pd}
\end{figure}
Using the logarithmic parametrization we find the RW endpoint to be a triple point independent of the quark masses, contrary to lattice results. First-order lines departing from the triple point are nicely visible.
Increasing the bare quark masses $m_0$ leads to larger effective quark masses. This results in growing ``RW legs'', see Fig.~\ref{fig:legsvarM0}. For $m_0$ larger than about $180$~MeV the first-order lines even reach across the $\mu=0$~axis. This scenario is shown in the right panel of Fig.~\ref{fig:legsvarM0} in comparison to the standard-parameter results.
If instead the coupling constant $g_S$ is increased, which likewise leads to larger constituent quark masses, the same effect is found~\cite{Morita}.
\begin{figure}
\includegraphics[height=0.26\textheight]{legsvarM0.eps}
\includegraphics[height=0.26\textheight]{pd2_M0-200.eps}
\caption{Left panel: ``RW legs'' in the $\theta-T$ phase diagram for different values of the bare quark mass $m_0$. Right panel: Phase diagram in the $\mu^2-T$ plane for two different values of the bare quark mass. Red solid (blue dashed) lines show first-order RW/deconfinement (deconfinement crossover) transitions for a high bare quark mass $m_0 = 200$ MeV. Thin lines show the RW, deconfinement and chiral transitions for the standard value of $m_0 = 5.5$ MeV. }
\label{fig:legsvarM0}
\end{figure}
Next we analyze the behavior of other Polyakov-loop potentials. Though all parametrizations are designed to reproduce pure-gauge lattice thermodynamics their effect on the RW endpoint is quite different.
The polynomial parametrization~\cite{RTW05} leads to a second-order transition for all examined quark masses. The reason is, that the polynomial parametrization shows a much weaker first-order transition in the heavy-quark limit.
\begin{figure}
\centering
\includegraphics[width=0.55\textwidth]{varb.eps}
\caption{Temperature and order of the RW transition as function of parameter~$b$.}
\label{fig:varb}
\end{figure}
Similarly, the Fukushima-type Polyakov-loop potential~\cite{Fukushima}, given by
\begin{equation*}
\frac{\mathcal{U}_\text{Fuku}}{T^4} = -b T \left( 54 e^{-a/T} \Phi\bar\Phi + \log\left[1 -6\Phi\bar\Phi + 4(\Phi^3+{\bar\Phi}^3) - 3(\Phi\bar\Phi)^2 \right] \right) \text,
\end{equation*}
produces a second-order transition for small quark masses and changes to first order only for very high quark masses where the PNJL model is not applicable any more. An alternative way to drive the system towards the pure gauge limit is to increase the global factor $b$ of the Fukushima-type Polyakov-loop potential. As presented in Fig.~\ref{fig:varb}, the RW endpoint changes from second to first order at about $b = 0.09 \Lambda^3$ whereas the default value for $N_f=2$ is $b=0.015 \Lambda^3$. For $b>0.5 \Lambda^3$ the ``RW legs'' reach across the temperature axis. With increasing $b$ the transition temperature approaches the heavy-quark limit of $T_c=270$~MeV.
We conclude, that the PNJL model together with currently available parametrizations for the Polyakov-loop potential is not able to reproduce the mass dependence found in lattice QCD studies. Sakai et.\,al.\ have shown that the 'entanglement' PNJL (EPNJL) model, which uses a Polyakov-loop dependent coupling $g_S$, reproduces the desired behavior~\cite{SakaiEPNJL}
\section{Summary}
We have shown that the choice of the Polyakov-loop potential pa\-ram\-e\-tri\-za\-tion has an important influence on the order of the RW phase transition endpoint. Modifying the strength of the quark degrees of freedom relative to the Polyakov-loop potential which models the gluon degrees of freedom, we find interesting changes in the phase structure at imaginary and real chemical potential. Results from lattice QCD should be used to constrain the Polyakov-loop potential parametrizations used in model studies.
\bigskip
The authors thank the organizers for an interesting workshop. D.\,S.\ acknowledges travel support by HIC for FAIR. This work was partially supported by the German Federal Ministry of Education and Research under project nr.\ 06DA9047I, the Helmholtz Alliance EMMI and the Helmholtz International Center for FAIR.
|
\section{Introduction}
\label{sec:introduction}
Geometric inequalities have an ancient history in Mathematics. A classical
example is the isoperimetric inequality for closed plane curves given by
\begin{equation}
\label{eq:54}
L^2 \geq 4\pi A,
\end{equation}
where $A$ is the area enclosed by a curve $C$ of length $L$, and where equality
holds if and only if $C$ is a circle (for a review on this subject see
\cite{Osserman78}). General Relativity is a geometric theory, hence it is not
surprising that geometric inequalities appear naturally in it. As we will see,
many of these inequalities are similar in spirit as the isoperimetric
inequality (\ref{eq:54}). However, General Relativity as a physical theory
provides an important extra ingredient. It is often the case that the
quantities involved have a clear physical interpretation and the expected
behavior of the gravitational and matter fields often suggest geometric
inequalities which can be highly non-trivial from the mathematical point of
view. The interplay between geometry and physics gives to geometric
inequalities in General Relativity their distinguished character.
A prominent example is the positive mass theorem. The physics suggests that the
mass of the spacetime (which is represented by a pure geometrical quantity
\cite{Arnowitt62}\cite{Bartnik86}\cite{chrusciel86}) should be positive and
equal to zero if and only if the spacetime is flat. From the geometrical mass
definition, without the physical picture, it would be very hard to conjecture
this inequality. In fact the proof turn out to be very subtle
\cite{Schoen79b}\cite{Schoen81}\cite{witten81}.
A key assumption in the positive mass theorem is that the matter fields should
satisfy an energy condition. This condition is expected to hold for all
physically realistic matter. It is remarkable that such a simple condition
encompass a huge class of physical models and that it translates into a pure
geometrical condition. This kind of general properties which do not depend
very much on the details of the model are not easy to find for astrophysical
objects (like stars or galaxies) which usually have a very complicated
structure. And hence it is difficult to obtain simple geometric inequalities
among the parameters that characterize them.
In contrast, black holes represent a unique class of very simple macroscopic
objects that play, in some sense, the role of `elementary particles' in the
theory. The black hole uniqueness theorem ensures that stationary black holes
in electro-vacuum are characterized by three parameters, which can be taken to
be the area $A$ of the black hole, the angular momentum $J$ and the charge
$q$. The mass $m$ is calculated in terms of these parameters by an explicit
formula (cf. equation (\ref{eq:11})). It is well known that these parameters
satisfy certain geometrical inequalities which restrict the range of them.
These inequalities are direct consequences of the explicit formula
(\ref{eq:11}). Among them, we note first the following
\begin{equation}
\label{eq:2bc}
m\geq \sqrt{\frac{A}{16\pi}},
\end{equation}
which will lead to the Penrose inequality for dynamical black holes. Also we
have the following two inequalities which will play a central role in this
article
\begin{equation}
\label{eq:24i}
m^2\geq \frac{q^2+\sqrt{q^4+4J^2}}{2}, \quad A \geq 4\pi \sqrt{q^4+4J^2} .
\end{equation}
The equality in (\ref{eq:2bc}) is achieved for the Schwarzschild black hole. The
equality in both inequalities \eqref{eq:24i} are achieved for extreme black
holes.
However black holes are not stationary in general. Astrophysical phenomena like
the formation of a black hole by gravitational collapse or a binary black hole
collision are highly dynamical. For such systems, the black hole can not be
characterized by few parameters as in the stationary case. In fact, even
stationary but non-vacuum black holes have a complicated structure (for example
black holes surrounded by a rotating ring of matter, see the numerical studies
in \cite{Ansorg05}). Remarkably, inequalities (\ref{eq:2bc})--(\ref{eq:24i})
extend (under appropriate assumptions) to the fully dynamical regime. Moreover,
inequalities (\ref{eq:2bc})--(\ref{eq:24i}) are deeply connected with
properties of the global evolution of Einstein equations, in particular with
the cosmic censorship conjecture. The main subject of this review is to present
a series of recent results which are mainly concerned with the dynamical
versions of inequalities (\ref{eq:24i}).
To extend the validity of inequalities (\ref{eq:2bc})--(\ref{eq:24i}) to
non-stationary black holes the first difficulty is how to define the physical
parameters involved, most notably the angular momentum $J$ of a dynamical black
hole. To define quasi-local quantities is in general a difficult problem (see
the review \cite{Szabados04}). However, for axially symmetric black holes, the
angular momentum (via Komar's formula) is well defined and it is conserved in
vacuum. Essentially for this reason inequalities (\ref{eq:24i}) have been
mostly studied for axially symmetric black holes. An exception are inequalities
which involves only the electric charge since the charge is well defined as a
quasi-local quantity without any symmetry assumption.
The plan of the article is the following. In section \ref{sec:physical-picture}
we describe the heuristic physical arguments that support these inequalities
and connect them with global properties of a gravitational collapse. In section
\ref{sec:results} we present an overview of the main results concerning these
inequalities that have been recently obtained. We also describe the main ideas
behind the proofs. Two important geometrical quantities involved in
(\ref{eq:24i}), mass and angular momentum, have distinguished properties in
axial symmetry (in contrast to the electric charge). These properties play a
fundamental role. We describe in some detail the angular momentum in section
\ref{sec:angul-moment-axial} and the mass in section
\ref{sec:mass-axial-symmetry}. Finally in section \ref{sec:open-problems} we
present the relevant open problem in this area.
\section{The physical picture}
\label{sec:physical-picture}
The most important example of a geometric inequality for dynamical black holes
is the Penrose inequality. In a seminal article Penrose \cite{Penrose73}
proposed a physical argument that connects global properties of the
gravitational collapse with geometric inequalities on the initial
conditions. For a recent review about this inequality see \cite{Mars:2009cj}
and references therein. Since it will play an important role in what follows,
let us review Penrose argument.
We will assume that the following statements hold in a gravitational collapse:
\begin{itemize}
\item[(i)] Gravitational collapse results in a black hole (weak cosmic
censorship).
\item[(ii)] The spacetime settles down to a stationary final
state. We will further assume that at some finite time all the matter have
fallen into the black hole and hence the exterior region is
electro-vacuum.
\end{itemize}
Conjectures (i) and (ii) constitute the standard picture of the gravitational
collapse. Relevant examples where this picture is confirmed (and where the role
of angular momentum is analyzed) are the collapse of neutron stars studied
numerically in \cite{Baiotti:2004wn} \cite{Giacomazzo:2011cv}.
Before going into the Penrose argument, let us analyze the final stationary
state postulated in (ii). The black hole uniqueness theorem implies that the
final state is given by the Kerr-Newman black hole (we emphasize however that
many important aspects of the black holes uniqueness still remain open, see
\cite{Chrusciel:2008js} for a recent review on this problem). Let us denote by
$m_0$, $A_0$, $J_0$ and $q_0$ the mass, area, angular momentum and charge of
the remainder Kerr-Newman black hole. In order to describe a black hole, the
parameters of the Kerr-Newman family of solutions of Einstein equations should
satisfy the remarkably inequality
\begin{equation}
\label{eq:13}
d\geq 0,
\end{equation}
where we have defined
\begin{equation}
\label{eq:8}
d= m_0^2-q_0^2-\frac{J_0^2}{m_0^2}.
\end{equation}
This inequality is equivalent to
\begin{equation}
\label{eq:24}
m_0^2\geq \frac{q_0^2+\sqrt{q_0^4+4J_0^2}}{2}.
\end{equation}
From Newtonian considerations (take $q=0$ for simplicity), we can interpret
this inequality as follows (see \cite{Wald71}): in a collapse the gravitational
attraction ($\approx m_0^2/r^2$) at the horizon ($r\approx m_0$) dominates over the
centrifugal repulsive forces ( $\approx J_0^2/m_0r^3$).
It is important to recall that the Kerr-Newman solution is well defined for any
choice of the parameters, but it only represents a black hole if the inequality
(\ref{eq:24}) is satisfied.
The Kerr-Newman black hole is called extreme if the equality in (\ref{eq:24})
is satisfied, namely
\begin{equation}
\label{eq:16}
m_0^2 = \frac{q_0^2+\sqrt{q_0^4+4J_0^2}}{2}.
\end{equation}
The area of the black hole horizon is given by the important formula
\begin{equation}
\label{eq:11}
A_0=4\pi \left(2m^2_0-q_0^2+2m_0 \sqrt{d} \right).
\end{equation}
Note that this expression has meaning only if the inequality (\ref{eq:24})
holds.
Penrose argument runs as follows. Let us consider a gravitational
collapse. Take a Cauchy surface $S$ in the spacetime such that the collapse has
already occurred. This is shown in figure \ref{fig:1}.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{penrose-ineq-2}
\caption{Schematic representation of a gravitational collapse.}
\label{fig:1}
\end{figure}
Let $\Sigma$ denotes the intersection of the event horizon with the Cauchy
surface $S$ and let $A$ be its area. Let $(m,q,J)$ be the total mass, charge
and angular momentum at spacelike infinity. These quantities can be computed
from the initial surface $S$. By the black hole area theorem we have that the
area of the black hole increase with time and hence
\begin{equation}
\label{eq:15}
A_0\geq A.
\end{equation}
Since gravitational waves carry positive energy, the total mass of the
spacetime should be bigger than the final mass of the black hole
\begin{equation}
\label{eq:4}
m\geq m_0.
\end{equation}
The difference $m-m_0$ is the total amount of gravitational radiation emitted
by the system.
The area $A_0$ of the remainder black hole is given by equation (\ref{eq:11})
in terms of the final parameters ($J_0, q_0, m_0$). It is a monotonically
increasing function of $m_0$ (for fixed $q_0$ and $J_0$), namely the derivative
$\partial A_0/\partial m_0$ is positive (we explicitly calculate this
derivative bellow). Then, using this monotonicity and inequalities
(\ref{eq:15}) and (\ref{eq:4}) we obtain
\begin{equation}
\label{eq:19c}
A\leq A_0 \leq 4\pi \left(2m^2-q_0^2+2m
\left(m^2-q_0^2-\frac{J_0^2}{m^2}\right)^{1/2} \right),
\end{equation}
where the important point is that in the right hand side appears the total mass
$m$ (instead of $m_0$), which can be calculated on the Cauchy surface $S$. The
parameters $q_0$ and $J_0$ are not known a priori but since they appear with
negative sign we have
\begin{equation}
\label{eq:20c}
A\leq A_0 \leq 4\pi \left(2m^2-q_0^2+2m
\left(m^2-q_0^2-\frac{J_0^2}{m^2}\right)^{1/2} \right)\leq 16\pi m^2.
\end{equation}
Remains still an important point: how to estimate the area $A$ of $\Sigma$ in
terms of geometrical quantities that can be locally computed on the initial
conditions. Recall that in order to know the location of the event horizon the
whole spacetime is needed. Assume that the surface $S$ contains a future
trapped two-surface $\Sigma_0$ (the trapped condition is a local property). By a
general result on black hole spacetimes we have that the surface $\Sigma_0$
should be contained in $\Sigma$. But that does not necessarily means that the
area of $\Sigma_0$ is smaller than the area of $\Sigma$. Consider all surfaces
$\tilde \Sigma$ enclosing $\Sigma_0$. Denote by $A_{\min}(\Sigma_0)$ the
infimum of the areas of all such surfaces. Then we clearly have that
$A(\Sigma)\geq A_{\min}(\Sigma_0)$. The advantage of this construction is that
$A_{\min}(\Sigma_0)$ is a quantity that can be computed from the Cauchy surface
$S$. Using this inequality and inequalities (\ref{eq:19c}) and (\ref{eq:20c})
we finally obtain the Penrose inequality
\begin{equation}
\label{eq:58}
m\geq \sqrt{\frac{ A_{\min}(\Sigma_0)}{16\pi}}.
\end{equation}
For further discussion we refer to \cite{Mars:2009cj} and references therein.
Penrose argument is remarkable because its end up in an inequality that can be
written purely in terms of the initial conditions. On the other hand, the proof
of such inequality gives indirect evidences of the validity of the conjectures
(i) and (ii).
Can we include the parameters $q$ and $J$ in the inequality
(\ref{eq:58}) to get a stronger version of it? The problem is, of course, how
to relate the final state parameters $(q_0, J_0)$ with the initial state ones
$(q,J)$.
If the matter fields are not charged, then the charge is conserved, namely
\begin{equation}
\label{eq:21}
q=q_0.
\end{equation}
And hence in that case we have the following version of the Penrose inequality
with charge
\begin{equation}
\label{eq:22}
A_{\min}(\Sigma_0)\leq A\leq 4\pi \left(2m^2-q^2+2m
\left(m^2-q^2\right)^{1/2} \right).
\end{equation}
The case of angular momentum is more complicated. Angular momentum is in
general non-conserved. There exists no simple relation between the total
angular momentum $J$ of the initial conditions and the angular momentum $J_0$
of the final black hole. For example, a system can have $J=0$ initially, but
collapse to a black hole with final angular momentum $J_0\neq 0$. We can
imagine that on the initial conditions there are two parts with opposite
angular momentum, one of them falls in to the black hole and the other scape to
infinity.
Axially symmetric vacuum spacetimes constitute a remarkable exception because
the angular momentum is conserved. In that case we have
\begin{equation}
\label{eq:59}
J=J_0.
\end{equation}
We discuss this conservation law in detail in section
\ref{sec:axially-symm-init}. The physical interpretation of (\ref{eq:59}) is
that axially symmetric gravitational waves do not carry angular momentum.
For non-vacuum axially symmetric spacetimes the angular momentum is no longer
conserved. Matter can transfer angular momentum even in axial symmetry. This
is also true for the electromagnetic field. However in the electro-vacuum case
a remarkably effect occurs. First, the sum of the gravitational and the
electromagnetic angular momentum is conserved. Second, at spacelike infinity
only the gravitational angular momentum is non-zero. In section
\ref{sec:angul-moment-axial} we prove this two facts. Hence if $J$ and $J_0$
denotes now the total angular momentum, then the conservation (\ref{eq:59})
still holds for axially symmetric electro-vacuum spacetimes and we obtain the
full Penrose inequality valid for axially symmetric electro-vacuum initial
conditions
\begin{equation}
\label{eq:23}
A_{\min}(\Sigma_0)\leq A\leq 4\pi \left(2m^2-q^2+2m
\left(m^2-q^2-\frac{J^2}{m^2}\right)^{1/2} \right).
\end{equation}
We emphasize that in this inequality the total angular momentum $J$ can be
computed at any closed surface that surround the black hole using the formula
(\ref{eqcm:21}). When this surface
is at infinity the angular momentum is given by the gravitational angular
momentum (i.e. the Komar integral of the axial Killing field).
Inequality (\ref{eq:23}) implies the bound
\begin{equation}
\label{eq:60}
m^2\geq \frac{q^2+\sqrt{q^4+4J^2}}{2}.
\end{equation}
Of course, inequality (\ref{eq:60}) can be deduced directly with the same
argument without using the area theorem. The first place where this conjecture
was formulated is in \cite{Friedman82} (see also \cite{Horowitz84}).
Inequality (\ref{eq:60}) can be viewed as a simplified version of the Penrose
inequality. The major difference is that the area of the horizon does not
appears in (\ref{eq:60}). Only charges, which are essentially topological,
appear in the right hand side of this inequality.
Inequality (\ref{eq:60}) is a global inequality for two reasons. First, it
involves the total mass $m$ of the spacetime. Second it assumes global
restrictions on the initial data: axial symmetry and electro-vacuum. We will
discuss these assumptions in more detail in section \ref{sec:results}.
The area $A$, the angular momentum $J$ in axial symmetry, and the charge $q$
are quasi-local quantities (in particular, the right hand side of (\ref{eq:60})
is purely quasi-local). Namely they carry information on a bounded region of
the spacetime. In contrast with a local quantity like a tensor field which
depends on a point of the spacetime or a global quantities (like the total
mass) which depends on the whole initial conditions. A natural question is
whether dynamical black holes satisfy purely quasi-local inequalities. The
relevance of this kind of inequalities is that they provide a much finer
control on the dynamics of a black holes than the global versions.
It is well known that the energy of the gravitational field can not be
represented by a local quantity. The best one can hope is to obtain a
quasi-local expression. These are the so called quasi-local mass definition
(see the review article \cite{Szabados04} and reference therein). Consider the
formula (\ref{eq:11}) for the horizon area for the Kerr-Newman black
holes. From this expression, we can write the mass in terms of the other
parameters as follows
\begin{equation}
\label{eq:32}
m_{bh}= \sqrt{\frac{A}{16\pi}+ \frac{q^2}{2}+\frac{\pi(q^4 +4J^2)}{A}}.
\end{equation}
In this equation we have dropped the subindice $0$ in the right hand side and
also denoted the mass in the left hand side by $m_{bh}$ to emphasize that this
expression can be in principle defined for any black hole (i.e. not necessarily
stationary). With this interpretation, this expression is known as the
Christodoulou \cite{Christodoulou70} mass of the black hole.
For a dynamical black hole the expression (\ref{eq:32}) is in principle just a
definition. Does the formula \eqref{eq:32} represents the quasi-local mass of
a non-stationary black hole? Let us analyze its physical behavior. We discuss
first the relation of this quasi-local mass with the total mass of the
spacetime. For only one black hole we expect the following inequality to
be true
\begin{equation}
\label{eq:53}
m \geq m_{bh}.
\end{equation}
This inequality implies the Penrose inequality (\ref{eq:23}) but it is stronger
(see the discussion in \cite{Mars:2009cj}). However, it is important to
emphasize that for the case of many black holes this inequality does not holds. In
fact it is possible to find counter examples if we take the area as additive
\cite{Weinstein05} or the quasi-local masses as additives
\cite{Dain:2010pj}. This is expected, since the interaction energy of the black
holes need to be taken into account (see the discussion in \cite{Dain:2010pj}).
For only one black hole, the inequality (\ref{eq:53}) in axial symmetry is an
open problem in this general form, we will further discuss it in section
\ref{sec:open-problems}.
We discuss now the purely quasi-local properties of (\ref{eq:32}).
The formula \eqref{eq:32} trivially satisfies the inequality (\ref{eq:60}).
This is, of course, just because the Kerr black hole satisfies this bound.
Hence, if we accept \eqref{eq:32} as the correct formula for the quasi-local
mass of an axially symmetric black hole, then (\ref{eq:32}) provides the,
rather trivial, quasi-local version of \eqref{eq:60}.
Consider the evolution of $m_{bh}$. By the area theorem, we know that the horizon
area will increase. If we assume axial symmetry and electro-vacuum, then the
total angular momentum (gravitational plus electromagnetic) will be conserved
at the quasi-local level. On physical grounds, one would expect that in this
situation the quasi-local mass of the black hole should increase with the area,
since there is no mechanism at the classical level to extract mass from the
black hole. In effect, the only way to extract mass from a black hole is by
extracting angular momentum through a Penrose process. But angular momentum
transfer is forbidden in electro-vacuum axial symmetry. Then, one would expect
that both the area $A$ and the quasi-local mass $m_{bh}$ should monotonically
increase with time.
Let us take a time derivative of $m_{bh}$ (denoted by a dot). To analyze this, it
is illustrative to write down the complete differential, namely the first law of
thermodynamics
\begin{equation}
\label{eq:55}
\delta m_{bh}= \frac{\kappa}{8 \pi} \delta A + \Omega_H \delta J + \Phi_H
\delta q,
\end{equation}
where
\begin{equation}
\label{eq:56}
\kappa= \frac{1}{4 m_{bh}} \left(1- \left(\frac{4\pi}{A}\right)^2 (q^4+4J^2)\right),
\quad \Omega_H= \frac{4\pi J}{Am_{bh}}, \quad \Phi_H= \frac{4\pi (m_{bh}+\sqrt{d})
q}{A},
\end{equation}
where $m_{bh}$ is given by (\ref{eq:32}) and $d$ (defined in equation
(\ref{eq:8})) is written in terms of $A$ and $J$ and $q$ as
\begin{equation}
\label{eq:17}
d= \frac{1}{m_{bh}^2}\left(\frac{A}{16\pi}\right)^2 \left(1-(q^4+4J^2)\left(\frac{4\pi}{A} \right)^2 \right)^2.
\end{equation}
In equation (\ref{eq:55}) we have followed the standard notation for the
formulation of the first law, we emphasize however that in our context this
equation is a trivial consequence of (\ref{eq:32}).
Under our assumptions, from the formula
\eqref{eq:32} we obtain
\begin{equation}
\label{eq:34}
\dot m_{bh} = \frac{\kappa }{8\pi}\dot A ,
\end{equation}
were we have used that the angular momentum $J$ and the charge $q$ are
conserved. Since, by the area theorem, we have
\begin{equation}
\label{eq:9}
\dot A \geq 0,
\end{equation}
the time derivative of $m_{bh}$ will be positive (and hence the mass $m_{bh}$ will
increase with the area) if and only if $\kappa \geq 0$, that is
\begin{equation}
\label{eq:5}
4\pi \sqrt{q^4+4J^2} \leq A.
\end{equation}
Then, it is natural to conjecture that \eqref{eq:5} should be satisfied for any
black hole in an axially symmetry. If the horizon violate
\eqref{eq:5} then in the evolution the area will increase but the mass $m_{bh}$
will decrease. This will indicate that the quantity $m_{bh}$ has not the desired
physical meaning. Also, a rigidity statement is expected. Namely, the equality
in \eqref{eq:5} is reached only by the extreme Kerr black hole given by the
formula
\begin{equation}
\label{eq:6}
A = 4\pi\left (\sqrt{q^4+4J^2 }\right).
\end{equation}
The final picture is that the size of the black hole is bounded from bellow by
the charge and angular momentum, and the minimal size is realized by the
extreme Kerr-Newman black hole. This inequality provides a remarkable
quasi-local measure of how far a dynamical black hole is from the extreme case,
namely an `extremality criteria' in the spirit of \cite{Booth:2007wu}, although
restricted only to axial symmetry. In the article \cite{Dain:2007pk} it has
been conjectured that, within axially symmetry, to prove the stability of a
nearly extreme black hole is perhaps simpler than a Schwarzschild black
hole. It is possible that this quasi-local extremality criteria will have
relevant applications in this context. Note also that the inequality
\eqref{eq:5} allows to define, at least formally, the positive surface gravity
density (or temperature) of a dynamical black hole by the formula (\ref{eq:56})
(see Refs. \cite{Ashtekar03} \cite{Ashtekar02} for a related discussion of the
first law in dynamical horizons).
If inequality \eqref{eq:5} is true, then we have a non trivial monotonic
quantity (in addition to the black hole area) $m_{bh}$ in electro-vacuum
\begin{equation}
\label{eq:10b}
\dot m_{bh} \geq 0.
\end{equation}
It is important to emphasize that the physical arguments presented above in
support of \eqref{eq:5} are certainly weaker in comparison with the ones behind
the Penrose inequalities (\ref{eq:58}), (\ref{eq:22}) and (\ref{eq:23}). A
counter example of any of these inequality will prove that the standard picture
of the gravitational collapse is wrong. On the other hand, a counter example of
\eqref{eq:5} will just prove that the quasi-local mass \eqref{eq:32} is not
appropriate to describe the evolution of a non-stationary black hole. One can
imagine other expressions for quasi-local mass, may be more involved, in axial
symmetry. On the contrary, reversing the argument, a proof of \eqref{eq:5}
will certainly suggest that the mass \eqref{eq:32} has physical meaning for
non-stationary black holes as a natural quasi-local mass (at least in axial
symmetry). Also, the inequality \eqref{eq:5} provide a non trivial control of
the size of a black hole valid at any time.
Finally, it is important to explore the physical scope of validity of these
geometrical inequalities. Are they valid for other macroscopic objects? The
Penrose inequality (\ref{eq:23}) is clearly not valid for an arbitrary region
in the spacetime. Namely, consider an arbitrary 2-surface of area $A$, which
is not necessarily a black hole boundary. We can make $A$ arbitrary large
keeping the mass (total or quasi-local) small (for example, take a region in
Minkowski).
For the inequalities (\ref{eq:60}) and (\ref{eq:5}) the situation is less
obvious. To have an intuitive idea of the order of magnitude involved it is
important to include the relevant constants in these inequalities. Let $G$ be
the gravitational constant and $c$ the speed of light. Then, these inequalities
are written as follows
\begin{equation}
\label{eq:4b}
m^2\geq \frac{1}{G}\frac{q^2+\sqrt{q^4+4J^2c^2}}{2}.
\end{equation}
and
\begin{equation}
\label{eq:7b}
4\pi \frac{G}{c^4}\sqrt{q^4+4 J^2 c^2} \leq A.
\end{equation}
For the reader's convenience we include the explicit values of the constants
(in centimeters, grams and seconds)
\begin{equation}
\label{eq:10c}
G =6.67 \times 10^{-8} \, cm^3 g^{-1} s^{-2},\quad
c = 3 \times 10^{10} \,cm\, s^{-1}.
\end{equation}
The values of the fundamental physical constants used in this section were
taken from \cite{mohr08}.
Let us analyze first the global inequality (\ref{eq:4b}). It is useful to split
it into the cases with zero charge and zero angular momentum
respectively, namely
\begin{equation}
\label{eq:8b}
\sqrt{G} \geq \frac{|q|}{m},
\end{equation}
and
\begin{equation}
\label{eq:14}
\frac{G}{c}\geq \frac{|J|}{m^2}.
\end{equation}
For the inequality (\ref{eq:8b}) consider an electron and a proton, for these
particles the quotient on the right hand side is given by
\begin{align}
\label{eq:qme}
\frac{|q_e|}{m_e} &=0.53 \times 10^{18} \, g^{-1/2} cm^{3/2}s^{-1},\\
\frac{|q_e|}{m_p} & =2.87 \times 10^{14} \, g^{-1/2} cm^{3/2}s^{-1}. \label{eq:qmp}
\end{align}
Since in these units we have $\sqrt{G} \approx 10^{-4}$, we see these particles
grossly violate the global inequality (\ref{eq:8b}). And hence ordinary charged
matter would violate it also (a similar discussion has been presented in
\cite{Gibbons82} and \cite{Horowitz84}).
For the angular momentum case (\ref{eq:14}) we can also consider an elementary
particle. In that case the angular momentum of a particle with spin $s$ (recall
that $s=1/2$ for the electron and the proton), is given by
\begin{equation}
\label{eq:13b}
J=\sqrt{s(s+1)}\hbar, \quad \hbar =1.05 \times 10^{-27} \, cm^2 s^{-1} g,
\end{equation}
where $\hbar$ is the Planck constant. For example, for the electron we have
\begin{equation}
\label{eq:17c}
\frac{|J|}{m^2_e}=1.12 \times 10^{27}\, cm^2 s^{-1} g^{-1}.
\end{equation}
Since
\begin{equation}
\label{eq:19d}
\frac{G}{c} =2.22 \times 10^{-18} \, cm^2 s^{-1} g^{-1},
\end{equation}
inequality (\ref{eq:14}) is also violated by several order of magnitude for
elementary particles. Instead of an elementary particle we can consider an
ordinary rotating object. It is clear that there exists ordinary object for
which $J/m^2 \approx 1$ (say a rigid sphere of mass $1 g$, radius $1 cm$, and
angular velocity $1 s^{-1}$) and hence inequality (\ref{eq:14}) is also
violated for ordinary rotating bodies.
We conclude that ordinary matter does not satisfies in general the global
inequality (\ref{eq:4b}). This inequality should be interpreted as a property
of electro-vacuum gravitational fields on complete regular initial conditions
where both the charge and the angular momentum are ``produced by the topology''
and not by matter sources (unless, of course, that they are inside a black hole
horizon). By ``produced by the topology'' we mean the following. The angular
momentum and the electric charge are defined as integral over closed two
dimensional surfaces. In electro-vacuum these integrals are conserved (we
discuss this in detail in section \ref{sec:angul-moment-axial}) and hence they
are zero if the topology of the initial conditions is trivial (i.e. $\mathbb{R}^3$). In
order to have non-trivial charges in electro-vacuum the initial conditions
should have some ``holes''. This non-trivial topology signals the presence of
a black hole.
For the quasi-local inequality (\ref{eq:7b}) it is also convenient to
distinguish between the cases with zero charge and zero angular momentum
respectively
\begin{equation}
\label{eq:9b}
A\geq \frac{G}{c^4}4\pi q^2,
\end{equation}
and
\begin{equation}
\label{eq:15b}
A\geq 8\pi \frac{G}{c^3} |J|.
\end{equation}
Since the charge is discrete, in
unit of $q_e$, it make sense to calculate the following characteristic radius
\begin{equation}
\label{eq:20}
r_0= \frac{q_e \,G^{1/2}}{c^2}= 1.38 \times 10^{-34}\, cm.
\end{equation}
We see that $r_0$ is one order of magnitude less than the Planck length
$l_p$ given by
\begin{equation}
\label{eq:1}
l_p = \left(\frac{G\hbar}{c^3}\right)^{1/2}= 1.6 \times 10^{-33} \,cm.
\end{equation}
If we assume that the particle or the macroscopic object has spherical shape we
can define the area radius $r$ by $A=4\pi r^2$. Then, the inequality
(\ref{eq:9b}) for a particle of charge $q_e$ has the form
\begin{equation}
\label{eq:21b}
r\geq r_0.
\end{equation}
The proton charge radius is $r_p = 0.87 \times 10^{-12} \, cm$ according to
\cite{pohl10} and $r_p = 0.84 \times 10^{-12} \, cm$ according to the recent
calculation presented in \cite{mohr08}. Hence, the proton satisfies inequality
(\ref{eq:21b}). This inequality is also consistent with the upper bound for the
electron radius $10^{-20} \, cm$ measured in \cite{dehmelt88}.
For the case of angular momentum, using the relation (\ref{eq:13b}) we can
compute the following quotient for an elementary particle
\begin{equation}
\label{eq:22b}
r_0=\left( 2\frac{G}{c^3}|J|\right)^{1/2}=\sqrt{2} (s(s+1))^{1/4} l_p.
\end{equation}
We see, that for a particle of spin $s$ of order $1$ the minimal radius is of
the order of the Planck length $l_p$ and hence it is also satisfied for
elementary particles.
More relevant is the case of an ordinary rotating
body. The angular momentum $J$ of a rigid body is given by
\begin{equation}
\label{eq:64m}
J=I \omega,
\end{equation}
where $I$ is the moment of inertia and $\omega$ the angular velocity.
Consider an ellipsoid of revolution with semi-axes $a$
and $b$, rotating along the $b$ axis. The moment of inertia along the axis of
rotation is given by
\begin{equation}
\label{eq:65}
I=\frac{2}{5}m a^2,
\end{equation}
where $m$ is the mass of the ellipsoid, which is assumed to have constant
density. The area of the ellipsoid satisfies the following elementary
inequality
\begin{equation}
\label{eq:87}
A \geq 2\pi a^2.
\end{equation}
The equality in (\ref{eq:87}) is achieved in the limit $b\to 0$, namely, the
bound is sharp.
Inserting equations (\ref{eq:64m}) and (\ref{eq:65}) in to the inequality
(\ref{eq:15b}), and using the bound (\ref{eq:87}) and the fact that it is sharp,
we obtain that the inequality (\ref{eq:15b}) is satisfied by the ellipsoid if
and only if the following inequality holds
\begin{equation}
\label{eq:50}
\frac{5}{8}\frac{c^3}{G}\geq m \omega.
\end{equation}
It is interesting to note that in the inequality (\ref{eq:50}) neither the area
nor the radii of the body appear. The inequality relates only the mass and the
angular velocity of the body. Also, the body is not assumed be nearly
spherical, the parameters for the ellipsoid are arbitrary.
The value of the left hand side of this inequality is
\begin{equation}
\label{eq:57}
\frac{5}{8}\frac{c^3}{G}= 2.53\times 10^{38}\, s^{-1} g.
\end{equation}
For the sun we have the following values
\begin{equation}
\label{eq:66h}
m_{sun}=1.989 \times 10^{33}\, g,\quad
\quad \omega= 2.90 \times 10^{-6} rad\, s^{-1},
\end{equation}
and hence
\begin{equation}
\label{eq:52}
m \omega =5.77 \times10^{27} \, s^{-1} g.
\end{equation}
We see that the inequality is satisfied for the sun. In order to violate
(\ref{eq:50}) a body should be very massive and highly spinning, a natural
candidate for that is a neutron star. For the fastest spinning neutron star
found to date (see \cite{Hessels:2006ze}) we have
\begin{equation}
\label{eq:66}
\omega \approx 4.5 \times 10^3 \, rad\, s^{-1}.
\end{equation}
Assuming that the neutron stars has about three solar masses (which appears to
be a reasonable upper bound for the mass, see \cite{Lattimer:2004pg}) we obtain
\begin{equation}
\label{eq:67}
m \omega\approx 2.7 \times 10^{37}\, s^{-1} g.
\end{equation}
The inequality (\ref{eq:50}) is still satisfied however the value (\ref{eq:67})
is remarkable close to the upper limit (\ref{eq:57}).
The example of the ellipsoid shows the elementary relation between shape and
angular momentum in classical mechanics which is valid even for non spherical
bodies. There is no such relation between shape and the electric charge of a
body. In fact, we will present some counter examples for charged highly
prolated objects that violate the inequality (\ref{eq:9b}). However,
remarkably enough, for charged `round surfaces' (we will define this concept
later on), an inequality between area and charge can be proved (see theorem
\ref{t:main3}). On the other hand, the example of the ellipsoid
suggests that the scope of validity of the inequality between area and angular
momentum (\ref{eq:15b}) (or the related inequality (\ref{eq:50})) for axial
symmetric bodies is much larger.
\section{Results and main ideas}
\label{sec:results}
In this section we present the main results concerning inequalities
(\ref{eq:60}) and (\ref{eq:5}) that haven been recently proved in the
literature. We also discuss the general strategy of their proofs.
\subsection{Global inequality}
\label{sec:global-inequality}
The proof of the inequality between total mass and charge (namely, setting
$J=0$ in (\ref{eq:60})), which is valid without any symmetry assumptions, has
been known for some time. The first proof was provided in \cite{Gibbons82} and
\cite{Gibbons83} using spinorial arguments similar to the Witten proof of the
positive mass theorem \cite{witten81}. See also \cite{Horowitz84}. A related
inequality was proved in \cite{Moreschi84} with similar techniques. In
\cite{Bartnik05} the proof was generalized to include low differentiable
metrics. For this inequality an interesting rigidity result is expected: the
equality holds if and only if the initial data are embed into the
Majumdar-Papapetrou static spacetime (see \cite{Hartle:1972ya} for a discussion
of these spacetimes in relations with black holes). The rigidity statement has
also been established, but with supplementary hypotheses, in \cite{Gibbons82}
and \cite{Chrusciel:2005ve}. Very recently a new proof was provided in
\cite{Khuri11} which removes all remaining hypothesis. Also in this article a
new approach is presented. The strategy is to combine the Jang equation method
with the spinorial proof of the positive mass theorem.
The inclusion of angular momentum in axial symmetry (which is the main subject
of this review) involves complete different techniques. In particular no
spinorial proof of these inequalities are available so far (see however
\cite{Zhang99} where a related inequality is proved using spinors). The first
proof of the global inequality (\ref{eq:60}) (with no electric charge) was
provided in a series of articles \cite{Dain05c}, \cite{Dain05d}, \cite{Dain05e}
which end up in the global proof given in \cite{Dain06c}.
In \cite{Chrusciel:2007dd} and \cite{Chrusciel:2007ak} the result was
generalized and the proof simplified. In \cite{Chrusciel:2009ki}
\cite{Costa:2009hn} the charge was included. As a sample of the most general
result currently available we present the following theorem proved in
\cite{Chrusciel:2009ki} and
\cite{Costa:2009hn}.
\begin{theorem}
\label{t:main-1}
Consider an axially symmetric, electro-vacuum, asymptotically flat and maximal
initial data set with two asymptotics ends. Let $m$, $J$ and $q$ denote the
total mass, angular momentum and charge respectively at one of the ends. Then,
the following inequality holds
\begin{equation}
\label{eq:42}
m^2\geq \frac{q^2+\sqrt{q^4+4J^2}}{2}.
\end{equation}
\end{theorem}
For the precise definition, fall off conditions an assumptions on the
electro-vacuum initial data we refer to \cite{Chrusciel:2009ki} and
\cite{Costa:2009hn}. For simplicity, in \ref{sec:axially-symm-init} we discuss
in detail only the pure vacuum case.
\begin{figure}
\centering
\includegraphics[width=4cm]{dos-ends-id}
\caption{Initial data with two asymptotically flat ends.}
\label{fig:4}
\end{figure}
Recall that for asymptotically flat initial data the total mass $m$, the total
charge $q$ and the total angular momentum $J$ (without any symmetry assumption)
are well defined as integrals over two-spheres at infinity for a given
asymptotic end. That is, all the quantities involved in (\ref{eq:42}) are well
defined for generic asymptotically flat data which are not necessarily axially
symmetric. However the inequality does not hold without the symmetry
assumption. General families of counter examples have been constructed in
\cite{huang11} for pure vacuum and complete manifolds.
Under the hypothesis of this theorem (namely, electro-vacuum and axial
symmetry) both the angular momentum and the electric charge are defined as
conserved quasi-local integrals (we discuss this in detail in section
\ref{sec:angul-moment-axial}). In particular, if the topology of the manifold
is trivial (i.e. $\mathbb{R}^3$), then these quantities are zero and hence theorem
\ref{t:main-1} reduces to the positive mass theorem. In order to have non-zero
charge or angular momentum we need to allow non-trivial topologies, for example
manifolds with two asymptotic ends as it is the case in theorem \ref{t:main-1}
(see figure \ref{fig:4}). An important initial data set that satisfies the
hypothesis of the theorem is provided by an slice $t=constant$ in the
non-extreme Kerr-Newman black hole in the standard Boyer-Lindquist coordinates.
This theorem has three main limitations: i) the initial data are assumed to be
maximal. ii) there is no rigidity statement. iii) the data are assumed to have
only two asymptotic ends. Let us discuss these points in more detail. The
maximal condition plays a crucial role in the proof since it ensure a positive
definite scalar curvature. A relevant open problem is how to remove this
condition, we will discuss it in more detail in section
\ref{sec:open-problems}.
Extreme Kerr-Newman initial data, which reach the equality in (\ref{eq:42}), is
not asymptotically flat in both ends. The data have a cylindrical end and an
asymptotically flat end (see figure \ref{fig:3}). Hence these data is excluded
in the hypothesis of theorem \ref{t:main-1}. In order to include the equality
case we need to enlarge the class of data. An example is given by the
following theorem proved in \cite{Dain06c} which includes the rigidity
statement.
\begin{theorem}
\label{t2}
Consider a vacuum Brill initial data set such that they satisfy condition 2.5
in \cite{Dain06c}. Then inequality
\begin{equation}
\label{eq:43}
m\geq \sqrt{|J|},
\end{equation}
holds. Moreover, the
equality in \eqref{eq:43} holds if and only if the data are a slice
of the extreme Kerr spacetime.
\end{theorem}
The precise definition of the Brill class of data can be seen in
\cite{Dain06c}. The main advantage of these kind of data is that they encompass
both class of asymptotics: cylindrical and asymptotic flatness. We discuss
this in section \ref{sec:axially-symm-init}. The condition 2.5 (see
\cite{Dain06c} for details) mentioned in this theorem implies that the initial
data have non trivial angular momentum only at one end, however multiple extra
ends with zero angular momentum are allowed. This condition involves also
other restrictions which are technical. In a very recent work \cite{Schoen11}
these technical conditions have been removed and also an interesting new
approach to the variational problem is presented.
\begin{figure}
\centering
\includegraphics[width=4cm]{cylindrical-end-initial-data}
\caption{The cylindrical end on extreme Kerr black hole initial data.}
\label{fig:3}
\end{figure}
The inequality (\ref{eq:42}) (which in particular implies (\ref{eq:43})) is
expected to hold for manifolds with an arbitrary number of asymptotic ends,
this generalization is probably to most important open problem regarding this
kind of inequalities (we discuss this in detail in the final section
\ref{sec:open-problems}). There exist, however, a very interesting partial
result \cite{Chrusciel:2007ak}. In order to describe it, we need to introduce
the mass functional $\mathcal{M}$, this functional is defined in section
\ref{sec:mass-axial-symmetry}. It plays a major role in all the proofs, as it
is explained in section \ref{sec:main-ideas}. This functional represents a
lower bound for the mass. Moreover, the global minimum of this functional
(under appropriate boundary conditions which preserve the angular momentum) is
achieved by an harmonic map with prescribed singularities. As we will see in
section \ref{sec:main-ideas} this is the main strategy in the proofs of all the
previous theorem which are valid for two asymptotic ends. Remarkably enough in
\cite{Chrusciel:2007ak} the existence and uniqueness of this singular harmonic
map has been proved also for manifolds with an arbitrary number of asymptotic
ends. In this article the following theorem is proved.
\begin{theorem}
\label{Piotr-Gilbert}
Consider an axially symmetric, vacuum asymptotically flat and maximal initial
data with $N$ asymptotic ends. Denote by $m_i$, $J_i$ ($i=1,\ldots N$) the
mass and angular momentum of the end $i$. Take an arbitrary end (say $1$),
then the mass at this end satisfies the inequality
\begin{equation}
m_1\geq \mathcal{M}(J_2,\ldots, J_N)
\end{equation}
where $ \mathcal{M}(J_2,\ldots, J_N)$ denotes the numerical value of the mass
functional $\mathcal{M}$ evaluated at the corresponding harmonic map.
\end{theorem}
This theorem reduces the proof of the inequality with multiples ends to compute
the value of the mass functional on the corresponding harmonic map and verify
the inequality
\begin{equation}
\label{eq:49}
\mathcal{M}(J_2,\ldots, J_N) \geq \sqrt{|J_1|}.
\end{equation}
We further discuss this theorem in section \ref{sec:global-inequalities-ideas}.
Strong numerical evidences that inequality (\ref{eq:49}) holds for three
asymptotic ends has been provided in \cite{Dain:2009qb}. The numerical methods
used in that article are related with the harmonic map structure of the
equations. We will describe them in section \ref{sec:harmonic-maps}.
All the previous results assume complete manifolds without inner
boundaries. The inclusion of inner boundary is important to prove the Penrose
inequality with angular momentum (\ref{eq:23}). Boundary conditions in
relation with the mass functional $\mathcal{M}$ where studied in \cite{Gibbons06} in
order to prove a version of Penrose inequality in axial symmetry. In
\cite{Chrusciel:2011eu} inner boundaries were also included and an interesting
new lower bound for the mass is obtained which depends only on the inner
boundary. Finally, we mention that in \cite{Jaramillo:2007mi} numerical
evidences for the validity of the Penrose inequality (\ref{eq:23}) has been
presented.
\subsection{Quasi-local inequalities}
\label{sec:quasi-local-ineq}
Quasi-local inequalities between area and charge has been proved in
\cite{Gibbons:1998zr} for stable minimal surfaces on time symmetric initial
data. The following theorem proved in \cite{Dain:2011kb} generalize this
result for generic dynamical black holes.
Consider Einstein equations with cosmological constant $\Lambda$
\begin{equation}
\label{eq:3aa}
G_{\mu\nu}=8\pi (T^{EM}_{\mu\nu}+T_{\mu\nu})-\Lambda g_{\mu\nu},
\end{equation}
where $T^{EM}_{\mu\nu}$ is the electromagnetic energy-momentum tensor defined
in terms of the electromagnetic field $F_{\mu\nu}$ by (\ref{eq:27}). The
electric charge of an arbitrary closed, oriented, two-surface $\mathcal{S}$ embedded in
the spacetime is defined by (\ref{eq:26}).
\begin{theorem}
\label{t:Aq}
Given a closed marginally trapped surface ${\cal S}$ satisfying spacetime
stably outermost condition, in a spacetime which satisfies Einstein equations
(\ref{eq:3aa}) with non-negative cosmological constant $\Lambda$ and such that
the non-electromagnetic matter fields $T_{\mu\nu}$ fulfill the dominant
energy condition, the following inequality holds:
\begin{equation}
\label{e:inequalityq}
A \geq 4\pi q^2 ,
\end{equation}
where $A$ and $q$ are the area and the charge of ${\cal S}$.
\end{theorem}
For the definition of marginally trapped surfaces (which is standard) and the
stably condition see \cite{Andersson:2005gq} \cite{andersson08}
\cite{Jaramillo:2011pg} \cite{Dain:2011kb} (see also \cite{Hayward93}
\cite{Racz:2008tf}). This theorem is a completely quasi-local result that
applies to general dynamical black holes without any symmetry assumption. It is
also important to emphasize that the matter is not assumed to be uncharged,
namely it is allowed that $\nabla_\mu F^{\mu\nu}\neq 0$ (which is equivalent to
$\nabla^\mu T^{EM}_{\mu\nu}\neq 0$ ). The only condition imposed in the
non-electromagnetic matter field stress-energy tensor $T_{\mu\nu}$ is that it
satisfies the dominant energy condition. In this theorem it is also possible to
include the magnetic charge and Yang-Mills charges (see \cite{Dain:2011kb}
and \cite{Jaram12}).
In \cite{Simon:2011zf} an interesting generalization of theorem \ref{t:Aq} is
presented in which the stability requirement is removed at the expense of
introducing the principal eigenvalue of the stability operator and also the
cosmological constant (with arbitrary sign) is added.
At the end of section \ref{sec:physical-picture} we have observed that a
variant of inequality (\ref{e:inequalityq}) is expected to hold for ordinary
macroscopic charged object (which are not necessarily black holes) at least if
they are `round enough'. In fact there exists an interesting and highly
non-trivial counter example to (\ref{e:inequalityq}) for macroscopic
objects. This counter example was constructed by W. Bonnor in \cite{bonnor98}
(see also the discussion in \cite{Dain:2011kb}) and it can be summarized as
follows: for any given positive number $k$, there exist static, isolated,
non-singular bodies, satisfying the energy conditions, whose surface area $A$
satisfies $A\leq k q^2$ . The body is a highly prolated spheroid of
electrically counterpoised dust. From the physical point of view we are saying
that for an ordinary charged object (in contrast to a black hole) we need to
control another parameter (the ‘roundness’) in order to obtain an inequality
between area and charge. Remarkably enough it is possible to encode this
intuition in the geometrical concept of isoperimetric surface: we say that a
surface $\mathcal{S}$ is isoperimetric (or `round') if among all surfaces that enclose
the same volume as $\mathcal{S}$ does, $\mathcal{S}$ has the least area. Then, based on the
results proved in \cite{christodoulou88} the following theorem was obtained in
\cite{Dain:2011kb} for isoperimetric surfaces.
\begin{theorem}
\label{t:main3}
Consider an electro-vacuum, maximal initial data, with a non-negative
cosmological constant. Assume that $\mathcal{S}$ is a stable isoperimetric sphere. Then
\begin{equation}
\label{eq:isoparametric}
A \geq \frac{4\pi}{3} q^{2},
\end{equation}
where $q$ is the electric charge of
$\mathcal{S}$.
\end{theorem}
Note that inequality (\ref{eq:isoparametric}) has a different coefficient as
(\ref{e:inequalityq}). We recall that the notion of stable isoperimetric surface is very
similar to the case of a stable minimal surface: the differential operator is
identical, the only difference is that the allowed test functions should
integrate to zero on the surface, this is precisely the condition that the
deformations preserve the volume (see \cite{Barbosa88}).
It is also possible to prove interesting variants of theorem \ref{t:Aq} which
are valid for generic surfaces (i.e. not necessarily trapped or minimal) but in
order to obtain these results global assumptions on the initial data should be
made (that is, in contrast to theorem \ref{t:Aq}, these are not a purely
quasi-local results): the two surfaces are embedded on initial conditions that
are complete, maximal and asymptotically flat. The non electromagnetic matter
fields are assumed to be non charged and they should satisfy the dominant
energy condition on the whole initial data (see Theorem 2.2 in
\cite{Dain:2011kb}).
As in the case of the global inequality, quasi-local inequalities with angular
momentum involve different techniques in comparison with the pure charged
case. Their study star very recently.
The quasi-local inequality with angular momentum and charge (\ref{eq:5}) was
first conjectured to hold in stationary spacetimes in \cite{Ansorg:2007fh}. In
that article the extreme limit of this inequality was analyzed and also
numerical evidences for the validity in the stationary case was presented
(using the numerical method and code developed in \cite{Ansorg05}). In a series
of articles \cite{hennig08} \cite{Hennig:2008zy} the inequality (\ref{eq:5})
was proved for stationary black holes. See also the review article
\cite{Ansorg:2010ru}.
It is important to emphasize that the stationary non-vacuum case is highly
non-trivial. The physical situation is, for example, a black hole surrounded by
a ring of matter (which do not touch the black hole). To illustrate the
complexity of this case we mention that the Komar mass (which is only defined
in the stationary case) can be negative for these black holes (for the
Kerr-Newman black hole is always positive), see \cite{Ansorg:2006qt}
\cite{Ansorg:2007vt}. It is interesting to mention that for this
class of stationary spacetimes there exists a remarkable relation of the form
$(8\pi J)^2+ (4\pi q^2)^2=A^+A^-$, where $A^+$ and $A^-$ denote the areas of
event and Cauchy horizon. This result have been proved in the following series
of articles \cite{Ansorg:2009yi} \cite{Hennig:2009aa} \cite{Ansorg:2008bv}.
In the dynamical regime, the inequality was conjectured to hold in
\cite{dain10d} based on the heuristic argument mentioned in section
\ref{sec:physical-picture}. In that article also the main relevant techniques
for its proof were introduced, namely the mass functional on the surface and
its connections with the area (we discuss this in section
\ref{sec:two-surfaces-mass}). A global proof (but with technical restrictions)
was obtained in \cite{Acena:2010ws} \cite{Clement:2011kz}. The first general
and pure quasi-local result was proven in \cite{Dain:2011pi}, where the
relevant role of the stability condition for minimal surfaces was pointed out:
\begin{theorem}
\label{t:main-2}
Consider an axisymmetric, vacuum and maximal initial data, with a non-negative
cosmological constant. Assume that the initial data contain
an orientable closed stable minimal axially symmetric surface
$\mathcal{S}$. Then
\begin{equation}
\label{desigualdad}
A \geq 8\pi |J|,
\end{equation}
where $A$ is the area and $J$ the angular momentum of $\mathcal{S}$. Moreover, if
the equality in \eqref{desigualdad} holds then $\Lambda=0$ and the local
geometry of the surface $\mathcal{S}$ is an extreme Kerr throat sphere.
\end{theorem}
The extreme throat sphere geometry, with angular momentum $J$, was defined in
\cite{dain10d} (see also \cite{Acena:2010ws} and \cite{Dain:2011pi}). This
surface capture the local geometry near the horizon of an extreme Kerr black
hole and it is defined as follows. The sphere is embedded in an initial
data with intrinsic metric given by
\begin{equation}
\label{eq:gamma0}
\gamma_0=4J^2e^{-\sigma_0}d\theta^2+ e^{\sigma_0}\sin^2\theta d\phi^2,
\end{equation}
where $\sigma_0$ is given by (\ref{datos}).
Moreover, the sphere must be totally geodesic, the twist
potential evaluated at the surface must be given by $\omega_0$ defined by (\ref{datos})
and the components of the second fundamental
\begin{equation}
K_{ij}\xi^i= K_{ij}n^j n^i= K_{ij}\eta^j \eta^i = 0,
\end{equation}
must vanish at the surface. Here $K_{ij}$ denotes the second fundamental form
of the initial data, $n^i$ the unit normal vector to the surface and $\eta^i$
the axial Killing field. Note that the functions $\sigma_0$ and $\omega_0$,
which characterize the intrinsic and extrinsic geometry of the surface
respectively, depend only on the angular momentum parameter $J$. The geometry
of axially symmetric initial data set are described in detail in section
\ref{sec:axially-symm-init}. In particular, the twist potential is determined
by the second fundamental form $K_{ij}$, using equation (\ref{eq:44}). The
adapted coordinates system used in (\ref{eq:gamma0}) is defined is section
\ref{s:potentials}.
This theorem has two main restriction: the first one is the maximal
condition. The second is vacuum. Remarkably enough, it is possible not only to
avoid both restrictions but also to provide a pure spacetime proof (that is, no
mention of a three-dimensional hypersurface) of this inequality, in which
axisymmetry is only imposed on $\mathcal{S}$. This generalization is proved in
\cite{Jaramillo:2011pg}:
\begin{theorem}
Given an axisymmetric closed marginally trapped surface $\mathcal{S}$ satisfying
the (axisymmetry-compatible) spacetime stably outermost condition, in a
spacetime with non-negative cosmological constant and fulfilling the dominant
energy condition, it holds the inequality
\begin{equation}
\label{e:inequality}
A \geq 8\pi |J|,
\end{equation}
where $A$ and $J$ are the area and (Komar) angular momentum of $\mathcal{S}$. If
equality holds, then $\mathcal{S}$ is a section of a non-expanding horizon with
the geometry of extreme Kerr throat sphere.
\end{theorem}
The concept of non-expanding horizon is explained in \cite{Jaramillo:2011pg},
it essentially means that the shear vanished at the surface.
It is important to note that the angular momentum that appears in
(\ref{e:inequality}) is the gravitational one (i.e. the Komar integral). The
matter fields have also angular momentum and it can be transferred to the black
hole, however the inequality (\ref{e:inequality}) remains true even in that
case. In fact this inequality is non-trivial for the Kerr-Newman black hole, we
discuss this in detail in section \ref{sec:angul-moment-axial}.
In \cite{reiris11} it has been pointed out that there exists a connexion
between the global inequalities described in section
\ref{sec:global-inequality} and the quasi-local inequalities. This is obtained
by linking the relevant mass functional $\mathcal{M}$ and $\mathcal{M}^{\Su}$ (we discuss these mass
functional in section \ref{sec:main-ideas}).
Inequality (\ref{e:inequality}) has played an important role in the proofs of
the non-existence of stationary two black holes configurations
(see \cite{Neugebauer:2011qb} \cite{Chrusciel:2011iv}).
Finally, we mention that there exists very interesting generalization of
(\ref{e:inequality}) to black holes in higher dimensions
\cite{Hollands:2011sy}.
\subsection{Main ideas}
\label{sec:main-ideas}
In this section we present the main ideas behind the proofs of the global
inequality (\ref{eq:42}) and the quasi-local inequality
(\ref{e:inequality}). This section should be consider as a guide in which the
technicalities are avoided. In section \ref{sec:angul-moment-axial} and
\ref{sec:mass-axial-symmetry} we discuss in details the main relevant
properties of the angular momentum and mass in axial symmetry which constitute
the essential part of the proofs.
\subsubsection{Global inequalities}
\label{sec:global-inequalities-ideas}
The starting point in the proof of the inequality (\ref{eq:42}) (we consider
the case $q=0$ for simplicity) is the formula for the total mass $m$ given by
Eq. (\ref{eqq:103}). This formula represents the total mass as a positive
definite integral over a maximal (here is where the maximal condition plays a
crucial role) initial surface. This integral representation is a generalization
of the Brill mass formula discovered in \cite{Brill59} (in section
\ref{sec:mass-axial-symmetry} we further discuss this formula and provide the
relevant references). The formula holds in a particular coordinate system which
is called isothermal (see lemma \ref{isothermal}).
The integrand in Eq. (\ref{eqq:103}) has two kind of terms: dynamical and
stationary. The dynamical terms vanished for an stationary solution like
Kerr. The stationary part lead to the relevant mass functional $\mathcal{M}$ defined by
(\ref{eqq:5c}). This functional provides an obvious lower bound to the total
mass, namely
\begin{equation}
\label{eq:2}
m\geq \mathcal{M}(\sigma, \omega).
\end{equation}
The mass functional $\mathcal{M}$ depends on two functions $\sigma$ and $\omega$. The
function $\sigma$ is essentially the norm of the axial Killing vector (see
equation (\ref{eq:83})). The function $\omega$ is the twist potential of the
Killing vector (see section \ref{sec:angul-moment-axial}). These two functions
can be freely prescribed on the initial data. This is an important and far from
obvious property since the constraint equations (\ref{const1})--(\ref{const2})
should be satisfied. This fact allows to formulate a variational principle for
the functional $\mathcal{M}$. The other important ingredient for this variational
principle is the behavior of the angular momentum. The angular momentum is
prescribed by the value of $\omega$ at the axis (see section
\ref{sec:angul-moment-axial}). Then, if the variations of $\omega$ vanishes at
the axis the angular momentum will be preserved. With this two ingredients, it
is hence possible to reduce the proof of the inequality (\ref{eq:42}) to a
pure variational problem for the functional $\mathcal{M}(\sigma, \omega)$.
The second step of the proof is to solve this variational problem. This is the
most difficult part and also the most interesting since it reveals the
geometric properties of the mass functional $\mathcal{M}$. Let us discuss the general
strategy of the proof.
Let $(\sigma_0, \omega_0)$ be the corresponding functions obtained from the
extreme Kerr initial data with angular momentum $J$, then we have that
\begin{equation}
\label{eq:3}
\mathcal{M}(\sigma_0, \omega_0)=\sqrt{|J|}.
\end{equation}
The heuristic discussed in section \ref{sec:physical-picture} and
Eq. (\ref{eq:10}) suggest that the following inequality holds
\begin{equation}
\label{eq:10}
\mathcal{M}(\sigma, \omega)\geq \sqrt{|J|},
\end{equation}
for all $(\sigma, \omega)$ such that $\omega$ has the same value at the axis as
the function $\omega_0$. Moreover, the equality in (\ref{eq:10}) is reached if
and only if $\sigma=\sigma_0$ and $\omega=\omega_0$. This is precisely the
variational problem. Note the variational problem is formulated purely in terms
of the functional $\mathcal{M}(\sigma, \omega)$, without any reference to the
constraint equations (\ref{const1})--(\ref{const2}).
The first evidence that this variational problem will have the expected
solution is that the Euler-Lagrange equations of the functional $\mathcal{M}$ are
equivalent to the stationary axially symmetric Einstein equations, in
particular extreme Kerr satisfies these equation \cite{Dain05c}. The second
evidence (which is harder to prove) is that the second variation of $\mathcal{M}$ is
positive definite evaluated at extreme Kerr \cite{Dain05d}. To prove this
positivity property it is crucial to make contact with the harmonic map theory
(in this case, trough the Carter identity). With these ingredients it is
possibly to show that extreme Kerr is a local minimum of the mass and hence the
inequality (\ref{eq:10}) is proved in an appropriate defined neighborhood of
extreme Kerr. This local proof was done in \cite{Dain05d}.
To have a global proof of (\ref{eq:10}) (i.e. without any smallness assumption)
more subtle properties of the mass functional are required. A crucial step is
to realize that the mass functional $\mathcal{M}$ is essentially the renormalized
energy of an harmonic map into the hyperbolic plane \cite{Dain06c} (we discuss
this in section \ref{sec:harmonic-maps}). This kind of harmonic maps have been
extensive studied in the literature. The problem here is that the map is
singular at the axis and hence the standard techniques do not apply directly.
To use the harmonic map theory we need to handle these singularities and this
is the main technical difficulty. In \cite{Dain06c} the proof of this
variational result was done using estimates, inspired in the work of
G. Weinstein \cite{Weinstein90} \cite{Weinstein92} \cite{Weinstein94}
\cite{Weinstein95} \cite{Weinstein96} \cite{Weinstein96b} (see also
\cite{Li92})), which rely on particular properties of this functional
(inversion symmetry). In subsequent works \cite{Chrusciel:2007dd},
\cite{Chrusciel:2007ak}, \cite{Chrusciel:2009ki}, \cite{Costa:2009hn} the proof
was simplified and improved using general results on harmonic maps (more
precisely the existence result \cite{Hildebrandt77}). In these proofs the
connection with the harmonic maps theory is more transparent and the problem of
the singularities is clearly isolated.
Finally let us mention the following important point regarding initial data
with multiples ends. The multiples ends appears in the variational problem
(\ref{eq:10}) as singular points of the functions $(\sigma, \omega)$.
Remarkably, even in that case it is possible to solve completely the
variational problem. This is precisely the content of theorem
\ref{Piotr-Gilbert} (proved in \cite{Chrusciel:2007ak}). The only missing piece
in order to prove the inequality (\ref{eq:42}) in that case is the
following. Theorem \ref{Piotr-Gilbert} ensure the existence of a global minimum
but the value of $\mathcal{M}$ at the global minimum is unknown. In the case of two
asymptotic ends we known that extreme Kerr is the global minimum and hence we
can explicitly compute the value (\ref{eq:3}).
\subsubsection{Quasi-local inequalities}
\label{sec:quasi-local-ineq-ideas}
The global inequality (\ref{eq:42}) applies to a complete three dimensional
manifold. In contrast the quasi-local inequality (\ref{e:inequality}) apply to
a closed two-surface. In principle it is a priori not clear at all that there
is a relation between these two kind of inequalities. The two physical
heuristic argument presented in section \ref{sec:physical-picture} in support
for them are very different. In particular, it is far from obvious that the
mass functional $\mathcal{M}$ can play a role for the quasi-local
inequalities. Remarkably enough, it turns out that a suitable adapted mass
functional $\mathcal{M}^{\Su}$ over a two-surface play a very similar role as $\mathcal{M}$. The
motivation for the definition of $\mathcal{M}^{\Su}$ given by (\ref{eqq:44}) is discussed in
section \ref{sec:two-surfaces-mass}. The new mass functional $\mathcal{M}^{\Su}$ and its
connection with the area represent the kernel of the proof of
(\ref{e:inequality}). Let us discuss this.
On the extreme Kerr initial data there exist an important canonical
two-surface, namely the intersection of the Cauchy surface with the horizon. On
a surface given by $t=constant$ in the Boyer-Lindquist coordinates this
two-surface is located at infinity on the cylindrical end (see figure
\ref{fig:3}). Let call $(\bar \sigma_0, \bar \omega_0)$ the value of the
functions $(\sigma_0, \omega_0)$ (defined in the previous section
\ref{sec:global-inequalities-ideas}) on this two-surface. The functions $(\bar
\sigma_0, \bar \omega_0)$ will play for $\mathcal{M}^{\Su}$ a similar role as the functions
$(\sigma_0, \omega_0)$ for $\mathcal{M}$. Namely, first they satisfy the Euler-Lagrange
equations of $\mathcal{M}^{\Su}$. Second, the second variation of $\mathcal{M}^{\Su}$ is positive
definite evaluated at $(\bar \sigma_0, \bar \omega_0)$ (see \cite{dain10d}).
There is also a connection with the harmonic maps energy (we discuss this in
section \ref{sec:two-surfaces-mass}). Using similar kind of arguments as
described in section \ref{sec:global-inequalities-ideas} it is possible to
prove the following inequality
\begin{equation}
\label{eq:12}
2|J|\leq e^{(\mathcal{M}^{\Su}(\sigma,\omega)-8)/8}
\end{equation}
for all functions $(\sigma, \omega)$ such that $\omega$ has the same values at
the poles of the two-surface as $\omega_0$. With equality if and only if we
have $\sigma=\bar \sigma_0$ and $\omega=\omega_0$. A local version of this
inequality was first proved in \cite{dain10d}. The global version (\ref{eq:12})
was proved in \cite{Acena:2010ws}. The rigidity statement was proved in
\cite{Dain:2011pi}. We emphasize that the inequality (\ref{eq:12}) (in complete
analogy to the inequality (\ref{eq:10})) is a property of the functional
$\mathcal{M}^{\Su}$, the geometry of the two-surface $\mathcal{S}$ does not intervene at all.
The inequality (\ref{eq:12}) is interesting but, in the light of the discussion
presented in \ref{sec:global-inequalities-ideas}, it is somehow expected. What
is crucial and completely unexpected is the relation between $\mathcal{M}^{\Su}$ and the
area of the surface $\mathcal{S}$. This relation was founded locally in \cite{dain10d}
and globally in \cite{Acena:2010ws}. By local in this context we mean that the
relation holds for surfaces in an appropriated defined neighborhood of an
extreme Kerr throat geometry. On the other hand, global means that the relation
holds for general surfaces. In \cite{Dain:2011pi} the important connection
with the stability condition was proved for minimal surfaces and in
\cite{Jaramillo:2011pg} for marginally trapped surfaces. The final inequality
essentially reads as follows. For a two-surface $\mathcal{S}$ such that: (i) it is
minimal \cite{Dain:2011pi} (or marginally trapped \cite{Jaramillo:2011pg}) and
(ii) it is stable, then the following inequality holds
\begin{equation}
\label{eq:18}
A\geq 4\pi e^{(\mathcal{M}^{\Su}-8)/8}.
\end{equation}
The minimal and marginally trapped conditions are requirements on the extrinsic
curvature of $\mathcal{S}$ (namely, on the trace of the second fundamental form). The
stability condition is a requirement on derivatives of the second fundamental
form. In the spacetime version presented in \cite{Jaramillo:2011pg}) the
inequality \eqref{eq:18} is a consequence of a flux inequality (see Lemma 1 in
that reference) where the geometric and physical meaning of each term is
apparent.
\section{Angular momentum in axial symmetry}
\label{sec:angul-moment-axial}
Axial symmetry plays a major role in the inequalities that include angular
momentum presented in the previous sections for two main reasons. For the
global inequality (\ref{eq:42}) is the conservation of angular momentum implied
by axial symmetry which is relevant. For the quasi-local inequality
(\ref{desigualdad}), is the very definition of quasi-local angular momentum
(only possible in axial symmetry) which is important. These two properties are
closed related for vacuum spacetimes, since the Komar integral provides both
the conservation law and the definition of quasi-local angular momentum. For
non-vacuum spacetimes in axial symmetry the Komar integral still provides a
meaningful expression for the gravitational angular momentum but it is no
longer conserved. Nevertheless, as we already mentioned in section
\ref{sec:physical-picture}, in the electro-vacuum case the sum of the
gravitational and electromagnetic angular is conserved, and hence in that case
is possible to prove the global inequality (\ref{eq:42}). In the general
non-vacuum case in axial symmetry (that is, with general matter sources which
are not electromagnetic) no simple and universal relation between mass and
angular momentum is expected. For example in \cite{Giacomazzo:2011cv} neutron
starts models in axial symmetry have been numerically constructed such that
they violate inequality (\ref{eq:42}). Nevertheless, remarkably, the
quasi-local inequality (\ref{desigualdad}) for black holes is still valid in
the general non-vacuum axially symmetric case.
In this section we summarize relevant results concerning angular momentum in
axial symmetry. Although these results are not new they are not easy to find in
the literature, notably, the conservation of angular momentum in the
electro-vacuum case and the relation between Komar integrals and potentials in
the presence of matter fields. The conservation of angular momentum in the
electro-vacuum, together with the relevant Komar integral for the
electromagnetic angular momentum, were discovered in \cite{Simon:1984qb}, in a
much general setting. We present a simpler derivation of this result in the
language of differential forms.
Let us begin with some general remarks about conserved quantities in General
Relativity (see also the review article \cite{jaramillo11c} for a related
discussion and \cite{Szabados04} for the general problem of how to define
quasi-local angular momentum without symmetries).
Let $M$ be a four dimensional manifold with metric $g_{\mu\nu}$ (with signature
$(-+++)$) and Levi-Civita connection $\nabla_\mu$. On this curved background,
let us consider an arbitrary energy-momentum tensor $T_{\mu\nu}$ which
satisfies the conservation equation
\begin{equation}
\label{eqcm:4}
\nabla_\mu T^{\mu\nu}=0.
\end{equation}
It is well known that if the spacetime admit a Killing vector field $\eta^\mu$
\begin{equation}
\label{eqcm:12}
\nabla_{(\mu} \eta_{\nu)}=0,
\end{equation}
then the vector
\begin{equation}
\label{eqcm:5}
K_{\mu}= T_{\mu \nu}\eta^\nu,
\end{equation}
is divergence free
\begin{equation}
\label{eq:64}
\nabla_\mu K^{\mu}=0.
\end{equation}
This equation provides an integral conservation law via Gauss theorem. For
some of the computations in this section it is convenient to use differential
forms instead of tensors. We will denote them with boldface. Let
$\boldsymbol{K}$ be the 1-form defined by (\ref{eqcm:5}). Equation
(\ref{eq:64}) is equivalent to
\begin{equation}
\label{eq:63}
d{}^*\boldsymbol{K}=0,
\end{equation}
where $d$ is the exterior derivative and the dual of a $p$ form is defined
with respect to the volume element $\epsilon_{\mu\nu\lambda\gamma}$ of the
metric $g_{\mu\nu}$ by the standard formula
\begin{equation}
\label{eq:61}
{}^*\alpha_{\mu_1\cdots \mu_{4-p}}=\frac{1}{p!}\alpha^{\nu_1\cdots \nu_p}
\epsilon_{\nu_1\cdots \nu_p \mu_1\cdots \mu_{4-p}}.
\end{equation}
Let $\Omega$ denotes a four-dimensional, orientable, region in $M$ and let
$\partial \Omega$ be its three-dimensional boundary. Then using (\ref{eq:63})
and the Stokes theorem we obtain
\begin{equation}
\label{eqcm:22}
0= \int_\Omega d{}^*\boldsymbol{K}= \int_{\partial \Omega} {}^*\boldsymbol{K}.
\end{equation}
Note that in this equation the region $\Omega$ is arbitrary and the boundary
$\partial \Omega$ can have many disconnected components.
Consider a spacelike three-surface $S$. The conserved quantity corresponding
to the Killing vector $\eta^\mu$ is defined with respect to $S$ by
\begin{equation}
\label{eqcm:8}
K(S)= \int_{S} {}^*\boldsymbol{K}.
\end{equation}
The interpretation of equation (\ref{eqcm:22}) in relation to the quantity
$K(S)$ is the following. Let $\Omega$ to be a timelike cylinder, such that its
boundary $\partial \Omega$ is formed by the bottom and the top spacelike
surfaces $S_1$ and $S_2$ and the timelike piece $\mathcal{C}$. Then we have
(taking the corresponding orientation)
\begin{equation}
\label{eqcm:7}
0= \int_{\partial \Omega} {}^*\boldsymbol{K}= K(S_1) - K(S_2) +
\int_{\mathcal{C}} {}^*\boldsymbol{K}.
\end{equation}
The integral over the timelike surface $\mathcal{C}$ is the flux of $K(S)$.
Equation (\ref{eqcm:7}) is interpreted as the conservation law of the quantity
$K(S)$. The region $\Omega$ can also be chosen to have a null boundary
$\mathcal{C}$, equation (\ref{eqcm:7}) remains identical and the interpretation
is similar.
If the Killing vector $\eta^\mu$ is also a symmetry of the tensor $T_{\mu\nu}$
(we have not assumed that so far), namely
\begin{equation}
\label{eqcm:10}
\pounds_\eta T_{\mu\nu}= 0,
\end{equation}
where $\pounds$ denote Lie derivatives, then the following vector is also
divergence free
\begin{equation}
\label{eqcm:11}
\hat K_\mu= 8\pi (T_{\mu \nu}\eta^\nu -\frac{1}{2} T \eta_{\mu}),
\end{equation}
where $T$ is the trace of $T_{\mu \nu}$. In fact there is a whole family of
divergence free tensors since $T\eta_\mu$ is divergence free. Hence the
previous discussion applies to this vector as well.
Note that the conserved quantities $K(S)$ are naturally defined as integrals
over spacelike three-surfaces. In flat spacetime it is possible to convert
these integrals into a boundary integral over two-surfaces. This gives the
quasi-local representation of conserved quantities (see the discussion in the
introduction of \cite{Szabados04}). In a curved background this is in general
not possible. However, as we will see, this possible for the particular case
of the electromagnetic field.
Before analyzing the angular momentum it is important to study the electric
charge. The electric charge is of course relevant in our discussion since it
appears in the inequalities discussed in the previous sections. But
more important, even if we want to analyze these inequalities in pure vacuum,
the electric charge represent the simpler `conserved charge' on a curved
spacetime. Its definition and properties serve as model for all the other
conserved quantities, like the angular momentum.
The Maxwell equations on $(M, g_{\mu\nu})$ are given by
\begin{align}
\label{eq:25}
\nabla^\mu F_{\mu\nu} & =-4\pi j_\nu, \\
\nabla_{[\mu} F_{\nu \alpha]} & =0.
\end{align}
In terms of forms, they are written as
\begin{align}
\label{eq:51}
d {}^* \mathbf{F} &=4\pi {}^* \mathbf{j},\\
d \mathbf{F} &=0.
\end{align}
The energy-momentum tensor of the electromagnetic field is given by
\begin{equation}
\label{eq:27}
T_{\mu\nu}=\frac{1}{4\pi}\left(F_{\mu\lambda}
F_\nu{}^{\lambda}-\frac{1}{4}g_{\mu\nu} F_{\lambda\gamma} F^{\lambda\gamma} \right).
\end{equation}
Let $\mathcal{S}$ be a closed orientable two-surface embedded in $M$ (in the
following, all two-surfaces will be assumed to be closed and orientable).
The electric charge $q$ of $\mathcal{S}$ is defined by
\begin{equation}
\label{eq:26}
q(\mathcal{S})=\frac{1}{4\pi}\int_{\mathcal{S}} {}^* \mathbf{F}.
\end{equation}
Let $S$ be a three-surface with boundary $\mathcal{S}$, then using Stokes theorem
and Maxwell equation (\ref{eq:51}) we obtain
\begin{equation}
\label{eq:28}
q(\mathcal{S})= \int_S {}^* \mathbf{j}.
\end{equation}
This equation is interpreted as follows.
From equation (\ref{eq:51}) we deduce the conservation law for the current
$\mathbf{j}$ analog to (\ref{eq:63}), namely
\begin{equation}
\label{eq:19}
d {}^* \mathbf{j}=0.
\end{equation}
And hence taking the same region $\Omega$ and using Stokes theorem we obtain
the analog expression as (\ref{eqcm:7}) for the current $\mathbf{j}$
\begin{equation}
\label{eq:29}
0= \int_{S_1} {}^* \mathbf{j}- \int_{S_2} {}^* \mathbf{j}+
\int_{\mathcal{C}} {}^* \mathbf{j}.
\end{equation}
Using (\ref{eq:28}) we finally obtain
\begin{equation}
\label{eq:30}
q(\mathcal{S}_1)- q(\mathcal{S}_2)= \int_{\mathcal{C}} {}^* \mathbf{j}.
\end{equation}
This is the conservation law for the electric charge. Note that in the left
hand side of (\ref{eq:30}) we have integrals over two-surfaces, in contrast
with (\ref{eqcm:7}) where integrals over three-surfaces appear. This is because
we have an extra equation (i.e. Maxwell equation (\ref{eq:51})) that allow us
to write the integral (\ref{eq:28}) in the form (\ref{eq:26}).
When $\mathbf{j}=0$ the charge has the same value, namely
\begin{equation}
\label{eq:62}
q(\mathcal{S}_1)= q(\mathcal{S}_2),
\end{equation}
and we say that the charge is strictly conserved.
We turn now to angular momentum for axially symmetric spacetimes. We begin with
the definition of axial symmetry.
\begin{definition}
\label{d:ax}
The spacetime $(M,g_{\mu\nu})$ is said to be axially symmetric if its group of
isometries has a subgroup isomorphic to $SO(2)$.
\end{definition}
We will denote by $\eta^\mu$ the Killing field generator of the axial
symmetry. The orbits of $\eta^\mu$ are either points or circles. The set of
point orbits $\Gamma$ is called the axis of symmetry. Assuming that $\Gamma$ is
a surface, it can be proved that $\eta^\mu$ is spacelike in a neighborhood of
$\Gamma$ (see \cite{Mars:1992cm}). We will further assume that the Killing
vector is always spacelike outside $\Gamma$. Note that if this condition is not
satisfied then the spacetime will have closed causal curved, in particular it
can not be globally hyperbolic.
The form $\eta_\mu$ will be denoted by $\boldsymbol{\eta}$, and the square of
its norm by $\eta$, namely
\begin{equation}
\label{eq:7}
\eta=\eta^\mu\eta_\mu=|\boldsymbol{\eta}|^2.
\end{equation}
We have used the notation $\eta^\mu$ to denote the Killing vector field and
$\eta$ to denote the square of its norm to be consistent with the
literature. However, in this section, to avoid confusions between $\eta^\mu$
and its square norm $\eta$, we will denote the vector field $\eta^\mu$ by
$\bar\eta$ in equations involving differential forms in the index free notation.
Consider now Einstein equations on an axially symmetric spacetime
\begin{equation}
\label{eqcm:2}
G_{\mu\nu}=8\pi T_{\mu\nu}.
\end{equation}
Note that the Killing equation (\ref{eqcm:12}) implies that $T_{\mu\nu}$
satisfies (\ref{eqcm:10}).
The Komar integral (it is also appropriate to call it the Komar charge) of the
Killing field is defined over a two-dimensional surface $\mathcal{S}$ as follows
\begin{equation}
\label{eqcm:1}
J(\mathcal{S})=\frac{1}{16\pi}\int_\mathcal{S}
\epsilon_{\mu\nu\lambda\gamma}\nabla^\lambda \eta^\gamma= \frac{1}{16\pi}
\int_\mathcal{S} {}^* d \boldsymbol{\eta}.
\end{equation}
Hence, as we discussed above, via Stokes theorem we obtain
\begin{equation}
\label{eqcm:52}
J(\mathcal{S})= \frac{1}{16\pi}
\int_S d {}^* d \boldsymbol{\eta},
\end{equation}
where $S$ is a three-dimensional surface with boundary $\mathcal{S}$.
It is a classical result \cite{Komar59} (see also \cite{Wald84}) that the
integrand in (\ref{eqcm:52}) can be computed in terms of the Ricci tensor
\begin{equation}
\label{eqcm:9}
\nabla_{[\mu} \left(\epsilon_{\nu\alpha] \beta\gamma}
\nabla^{\beta}\eta^{\gamma}\right)=\frac{2}{3}R^{\mu}{}_\nu\eta^\nu
\epsilon_{\mu\alpha \beta\gamma}.
\end{equation}
In terms of forms this equation is written as
\begin{equation}
\label{eqcm:13}
d {}^*d \boldsymbol{\eta}= 2 {}^* \boldsymbol{K},
\end{equation}
where we have defined the 1-form $\boldsymbol{K}$ by
\begin{equation}
\label{eqcm:14}
\boldsymbol{K}\equiv K_\mu =R_{\mu\nu}\eta^\nu.
\end{equation}
Using Einstein equations (\ref{eqcm:2}) the form $\boldsymbol{K}$ can be
written in terms of the energy momentum tensor
\begin{equation}
\label{eqcm:15}
K_\mu =8\pi (T_{\mu\nu}\eta^\nu-\frac{1}{2}T \eta_\mu) .
\end{equation}
Note that this expression is identical to (\ref{eqcm:11}). Then, repeating the
same argument, we obtain the conservation law for angular momentum in axial
symmetry which is the exact analog to the charge conservation (\ref{eq:30})
\begin{equation}
\label{eqcm:41}
J(\mathcal{S}_1)- J(\mathcal{S}_2)=\frac{1}{8\pi} \int_{\mathcal{C}} {}^* \mathbf{K}.
\end{equation}
The right hand side of this equation represent the change in the angular
momentum of the gravitational field which is produced by the left hand side,
namely the angular momentum of the matter fields. Note that the angular
momentum of the matter fields are written as integrals over three-dimensional
surfaces. In particular in vacuum we have the strict conservation of angular momentum
\begin{equation}
\label{eq:68}
J(\mathcal{S}_1)=J(\mathcal{S}_2).
\end{equation}
We consider now the case where the energy-momentum tensor in Einstein equations
(\ref{eqcm:2}) is given purely by the electromagnetic field (\ref{eq:27}). For
simplicity we consider the case with no currents $j=0$, which is the relevant
one since only in that case we get conserved quantities. In that case we have
that (\ref{eq:27}) satisfies the equation (\ref{eqcm:4}) and the source-free
Maxwell equations are given by
\begin{align}
\label{eqcm:51}
d {}^* \boldsymbol{F} &= 0,\\
d \boldsymbol{F} &=0. \label{eqcm:51b}
\end{align}
We also assume that the Maxwell
fields are axially symmetric, namely
\begin{equation}
\label{eqcm:29}
\pounds_\eta \boldsymbol{F}=0.
\end{equation}
Consider the 1-forms defined by
\begin{equation}
\label{eqcm:30}
\boldsymbol{\alpha}=\boldsymbol{F}(\bar\eta) , \quad \boldsymbol{\beta}={}^*
\boldsymbol{F}(\bar\eta),
\end{equation}
where we have used the standard notation
$\boldsymbol{F}(\bar\eta)=F_{\mu\nu}\eta^\mu$ to denote contractions of forms with
vector fields. Using the general expression for the action of the Lie
derivative on forms
\begin{equation}
\label{eqcm:42}
\pounds_{\bar\eta} \boldsymbol{\omega} =d [ \boldsymbol{\omega}(\bar\eta)]+ (d
\boldsymbol{\omega})(\bar\eta),
\end{equation}
Maxwell equations (\ref{eqcm:51})--(\ref{eqcm:51b}) and the condition
(\ref{eqcm:29}) we obtain
\begin{align}
\label{eqcm:31}
d \boldsymbol{\alpha} = 0, \quad d \boldsymbol{\beta}=0.
\end{align}
It follows that there exist locally functions $\chi$ and $\psi$ such that
\begin{equation}
\label{eqcm:33}
\boldsymbol{\alpha}=d \chi, \quad \boldsymbol{\beta}= d \psi.
\end{equation}
The form $\boldsymbol{K}$ defined by \eqref{eqcm:15} has the following
expression for the electromagnetic field
\begin{equation}
\label{eqcm:16}
\boldsymbol{K} =2\left(\boldsymbol{F}(\alpha)
-\frac{1}{4}\boldsymbol{\eta}|F|^2\right),
\end{equation}
We have that (see \cite{Weinstein96})
\begin{equation}
\label{eqcm:17}
{}^*(\boldsymbol{\eta} \wedge (\boldsymbol{F}(\alpha))=\boldsymbol{\alpha}
\wedge
\boldsymbol{ \beta}.
\end{equation}
Using (\ref{eqcm:33}) we obtain
\begin{equation}
\label{eqcm:18}
d (\boldsymbol{\alpha} \wedge \boldsymbol{ \beta})=0,
\end{equation}
and hence there exist locally a 1-form $\boldsymbol{\gamma}$ such that
\begin{equation}
\label{eqcm:19}
d\boldsymbol{\gamma} =\boldsymbol{\alpha} \wedge
\boldsymbol{ \beta},
\end{equation}
where $\gamma$ is given by
\begin{equation}
\label{eqcm:20}
\boldsymbol{\gamma}= \frac{1}{2}\left(\chi d \psi -\psi d \chi\right).
\end{equation}
Note that
\begin{equation}
\label{eqcm:32}
\boldsymbol{\gamma}(\bar\eta)=0.
\end{equation}
From equation (\ref{eqcm:17}) we deduce
\begin{equation}
\label{eqcm:35}
{}^*(\boldsymbol{\eta} \wedge \boldsymbol{K})=2 \boldsymbol{\alpha} \wedge
\boldsymbol{ \beta}.
\end{equation}
We use the following identity valid for arbitrary 1-forms
\begin{equation}
\label{eqcm:36}
{}^*(\boldsymbol{\eta} \wedge \boldsymbol{K})\wedge \boldsymbol{\eta}=\eta
{}^*\boldsymbol{K} - {}^*\boldsymbol{\eta} ( \boldsymbol{K}(\bar\eta)).
\end{equation}
Using (\ref{eqcm:19}) we finally obtain our main formula
\begin{equation}
\label{eqcm:37}
{}^*\boldsymbol{K}= 2d(\boldsymbol{\gamma} \wedge \boldsymbol{\hat\eta}) +
2\boldsymbol{\gamma} \wedge d \boldsymbol{\hat\eta} + {}^*\boldsymbol{\hat\eta} (
\boldsymbol{K} \cdot \boldsymbol{\eta}),
\end{equation}
where we have defined
\begin{equation}
\label{eqcm:38}
\boldsymbol{\hat\eta}= \frac{\boldsymbol{\eta}}{\eta}.
\end{equation}
It is important to note that
\begin{equation}
\label{eqcm:39}
d \boldsymbol{\hat\eta}(\bar\eta)=0.
\end{equation}
Using equation (\ref{eqcm:37}), we integrate ${}^*\boldsymbol{K}$ over a
three-surface $S$ tangential to $\eta^\mu$, with boundary $\mathcal{S}$. Using that
$\eta^\mu$ is tangential to $S$ it follows that the restriction
of the 3-form ${}^*\boldsymbol{\hat\eta}$ to $S$ is zero. For the second term in
(\ref{eqcm:37}) we use equations \eqref{eqcm:32} and \eqref{eqcm:39} to obtain
the same conclusion. Hence, we have
\begin{equation}
\label{eqcm:40}
\int_S {}^*\boldsymbol{K}= \int_S 2d(\boldsymbol{\gamma} \wedge
\boldsymbol{\hat\eta})= 2 \int_\mathcal{S} \boldsymbol{\gamma} \wedge
\boldsymbol{\hat\eta},
\end{equation}
where in the last equality we have used Stokes theorem.
We summarize the previous calculation in the following lemma, which is a
re-writing of the result that have been obtained in \cite{Simon:1984qb}.
\begin{lemma}
\label{l:ang-em}
Consider an axially symmetric spacetime for which the Einstein-Maxwell
equations (\ref{eqcm:2}), (\ref{eq:27}), (\ref{eqcm:51}), (\ref{eqcm:51b})
are satisfied. Let $S$ be an orientable three-surface, tangent to the axial
Killing field $\eta^\mu$, with boundary (possible disconnected)
$\mathcal{S}$. Then we have
\begin{equation}
\label{eqcm:23}
J(\mathcal{S})=\frac{1}{4\pi} \int_\mathcal{S} \boldsymbol{\gamma} \wedge
\boldsymbol{\hat\eta},
\end{equation}
where $J(\mathcal{S})$ is the Komar integral given by (\ref{eqcm:1}), $
\boldsymbol{\gamma}$ is defined in terms of the electromagnetic field by
(\ref{eqcm:20}) and $\boldsymbol{\hat\eta}$ is given by (\ref{eqcm:38}).
\end{lemma}
We can define a `total angular momentum' which is conserved in electro-vacuum,
namely
\begin{equation}
\label{eqcm:21}
J_T(\mathcal{S})= \frac{1}{16\pi} \int_\mathcal{S} {}^*d \boldsymbol \eta -4\,
\boldsymbol{\gamma}
\wedge \boldsymbol{\hat\eta} .
\end{equation}
We note that since the surface $\mathcal{S}$ is tangent to $\eta^\mu$ and all
the fields are axially symmetric, then the surface integrals are in fact line
integrals on the quotient manifold $M\setminus SO(2)$.
Formula (\ref{eqcm:21}) was studied in
\cite{Ashtekar:2001is}\cite{Ashtekar:2000sz} for rotating isolated horizons.
This formula has been also recently studied at null infinity in
connection with the center of mass and general definition of angular momentum
for asymptotically flat (at null infinity) spacetimes, see \cite{kozameh11}.
A very important example where lemma \ref{l:ang-em} applies is the Kerr-Newman
black hole. Consider the Kerr-Newman black hole with parameters $(m, a,
q)$. The total angular momentum is given by $J_T=am$. This is equal to the
Komar integral evaluated at infinity, since the electromagnetic field decay and
does not contribute at infinity. However at the horizon the Komar angular
momentum is not $J_T$. The angular momentum at the horizon has the
decomposition (\ref{eqcm:21}).
For the Kerr-Newman black hole the Komar angular momentum at the horizon $J$
is given by (see, for example, \cite{Poisson} page 222).
\begin{equation}
\label{eqk:1}
J= a \frac{r^2_+ +a^2}{2r_+} \left(1+\frac{q^2}{2a^2}\left(1- \frac{r^2_+
+a^2}{ar_+}\right) \arctan\left(\frac{a}{r_+}\right) \right),
\end{equation}
where $r_+$ is the horizon radius
\begin{equation}
\label{eqk:2}
r_+=m+(m^2-a^2-q^2 )^{1/2}.
\end{equation}
The area of the horizon is given by
\begin{equation}
\label{eqk:3}
A=4\pi(r^2_+ +a^2).
\end{equation}
Equation \eqref{eqk:3} is of course identical to equation \eqref{eq:11}.
As we already mentioned, it is well known that the area satisfy the inequality
\begin{equation}
\label{eqk:4}
A \geq 4\pi \sqrt{q^4+4J_T^2}.
\end{equation}
Which in particular implies
\begin{equation}
\label{eqk:5}
A \geq 8\pi |J_T|.
\end{equation}
But this inequality relates the total angular momentum $J_T$. It is a priori
not obvious if the following inequality holds
\begin{equation}
\label{eqk:6}
A \geq 8\pi |J|,
\end{equation}
where $J$ is given by (\ref{eqk:1}), namely the Komar integral at the horizon.
Note that this is precisely the inequality proved in theorem \ref{t:main-2} and
that the Kerr-Newman black hole satisfies all the hypothesis of that theorem.
Let us check explicitly that indeed (\ref{eqk:6}) is satisfied. The expression
(\ref{eqk:1}) is remarkably complicated, to better analyze it let us rewrite
it in the following form. Using (\ref{eqk:3}) we have
\begin{equation}
\label{eqk:7}
J=\frac{A}{8\pi}\epsilon,
\end{equation}
where
\begin{equation}
\label{eqk:8}
\epsilon=\frac{a}{r_+}\left(1+\frac{q^2}{2a^2}\left(1- \frac{r^2_+
+a^2}{ar_+}\right) \arctan\left(\frac{a}{r_+}\right) \right),
\end{equation}
Instead of using $a$ it is convenient to use $x=a/r_+$ as free parameter. In
terms of $x$ the function $\epsilon$ is written us
\begin{equation}
\label{eqk:9}
\epsilon=x\left( 1+ \frac{q^2}{2r^2_+} f(x)\right),
\end{equation}
where
\begin{equation}
\label{eqk:10}
f(x)=\frac{1}{x^2} \left(1- \left(\frac{1}{x}+x \right)\arctan(x)\right).
\end{equation}
We take as free parameters $(q,r_+,x)$. Note that $-1\leq x\leq 1$ and hence we have
\begin{equation}
\label{eqk:11}
|\epsilon| \leq \left| 1+ \frac{q^2}{2r^2_+} f(x)\right|.
\end{equation}
Fix $(q,r_+)$. It can be explicitly check that function $f(x)$
is non-positive and have a unique global minimum at $x=0$ where
$f(0)=-2/3$. Hence it follows that
\begin{equation}
\label{eqk:12}
|\epsilon| \leq 1.
\end{equation}
\subsection{Potentials}
\label{s:potentials}
The potentials for the axial Killing vector plays an important role in the mass
functional described in section \ref{sec:mass-axial-symmetry}.
It is instructive to analyze first the electric charge and its potential in
axial symmetry. Assume first that the Maxwell equations are source free. Then
we have found the potentials $\chi$ and $\psi$ defined by equation
(\ref{eqcm:33}). In particular, the potential $\psi$ determine the electric charge
over an axially symmetric two-surface $\mathcal{S}$. In order to see that it is
convenient to consider a tetrad $(l^\mu,k^\mu,\xi^\mu,\eta^\mu)$ and coordinate
system $(\theta,\phi)$ adapted to an axially symmetric two-surface defined as
follows (see figure \ref{fig:2}). For simplicity we will assume that $\mathcal{S}$ has
the topology of a two-sphere. Let us consider null vectors $\ell^\mu$ and
$k^\nu$ spanning the normal plane to $\mathcal{S}$ and normalized as $\ell^\mu k_\mu =
-1$, leaving a (boost) rescaling freedom $\ell'^\mu =f \ell^\mu$, $k'^\mu =
f^{-1} k^\mu$. By assumption $\eta^\mu$ is tangent to $\mathcal{S}$, it has on the
surface closed integral curves and vanishes exactly at two points which are the
intersection of the axis $\Gamma$ with $\mathcal{S}$. We normalize vector $\eta^\mu$ so
that its integral curves have an affine length of $2\pi$. Let us chose a
coordinate $\phi$ on $\mathcal{S}$ such that $\eta^\mu=\partial/\partial \phi$. The
other vector of the tetrad which is tangent to $\mathcal{S}$ and orthogonal to
$\eta^\mu$ will be denoted by $\xi^\mu$ and assume that it has unit norm. We
define the coordinate $\theta$ such that $\xi^\mu$ is proportional to
$\partial/\partial \theta$ and such that $\theta=\pi,0$ are the poles of $\mathcal{S}$.
The induced metric and the volume element on $\mathcal{S}$ (written as spacetime
projectors) are given by $q_{\mu\nu}=g_{\mu\nu}+\ell_\mu k_\nu+\ell_\nu k_\mu$
and $\epsilon_{\mu\nu}=2^{-1}\epsilon_{\lambda\gamma\mu\nu}\ell^\lambda k^\gamma$
respectively. The area measure on $\mathcal{S}$ is denoted by $ds$. Since the surface
is axially symmetric we have ${\cal L}_\eta q_{\mu\nu}=0$.
\begin{figure}
\centering
\includegraphics[width=3cm]{axial-two-surface-mu}
\caption{Adapted tetrad for an axially symmetric two-sphere}
\label{fig:2}
\end{figure}
Using this tetrad, the charge over an axially symmetric surface $\mathcal{S}$ is
written as follows
\begin{equation}
\label{eq:37}
q(\mathcal{S})=\frac{1}{2\pi}\int_\mathcal{S} \eta^{-1/2} \xi^\mu\beta_\mu \, ds.
\end{equation}
where $\beta_\mu$ is defined by (\ref{eqcm:30}). In the source free case we can
use the potential $\psi$ defined by (\ref{eqcm:33}) to obtain
\begin{equation}
\label{eq:36}
q(\mathcal{S})=\frac{1}{2\pi} \int_\mathcal{S} \eta^{-1/2} \xi^\mu \nabla_\mu \psi \, ds =
\int_0^\pi \partial_\theta \psi \, d\theta.
=\psi(\pi)-\psi(0),
\end{equation}
That is, the charge is given by the difference of the value of the potential
$\psi$ at the poles of the surface $\mathcal{S}$.
We have seen that the potential $\psi$ is only defined in the source-free case
as scalar function in the spacetime. However, on the surface $\mathcal{S}$ it is
always possible to define a potential $\bar \psi$ by the equation
\begin{equation}
\label{eq:39}
\xi^\mu\psi_\mu = \xi^\mu \nabla_\mu \bar \psi
\end{equation}
on the surface, since this equation involves only a derivative with respect to
$\theta$. The potential $\bar\psi$ is only defined on the surface. By
definition, in the source-free case where the potential $\psi$ also exists, the
two functions are equal up to a constants on the surface. This constant is
irrelevant since it does not affect the charge.
Consider now the potential for the angular momentum.
The twist vector of $\eta^\mu$ is defined by
\begin{equation}
\label{eqt:2}
\omega_\mu= \epsilon_{\mu\nu\lambda \gamma }\eta^\nu \nabla^\lambda\eta^\gamma.
\end{equation}
It is a well known result that the vacuum equations $R_{\mu\nu}=0$ imply that
\begin{equation}
\label{eqt:3}
d \boldsymbol{\omega}=0,
\end{equation}
where $\boldsymbol{\omega}$ is the 1-form defined by (\ref{eqt:2}). Hence
there exist a potential $\omega$ such that
\begin{equation}
\label{eq:31}
\boldsymbol{\omega}=d \omega.
\end{equation}
The function $\omega$ is the twist potential of the Killing vector $\eta^\mu$,
it contains all the information of the angular momentum as we will see. In the
non-vacuum case, the twist potential is not defined. In the following we will
not assume vacuum.
In terms of the adapted tetrad we have that the Komar expression is given by
\begin{equation}
\label{eqt:8}
J=\frac{1}{8\pi}\int_{\cal S} \nabla^{\mu} \eta^\nu l_{\mu}k_\nu \, ds.
\end{equation}
We the relation
\begin{equation}
\label{eqt:4}
\nabla_\mu \eta_\nu=\frac{1}{2}\eta^{-1} \epsilon_{\mu\nu \lambda \gamma}
\eta^\lambda \omega^\gamma+ \eta^{-1} \eta_{[\nu} \nabla_{\mu]}\eta.
\end{equation}
to write the Komar integral in the following form
\begin{equation}
\label{eqt:5}
J=\frac{1}{16\pi}\int_{\cal S} \eta^{-1} \epsilon_{\mu\nu\lambda \gamma } \eta^\lambda
\omega^\gamma \ell^{\mu} k^\nu \, ds.
\end{equation}
It is clear that the vector defined by
\begin{equation}
\label{eqt:6}
\xi_\gamma= \eta^{-1/2} \epsilon_{\mu\nu\lambda \gamma } \eta^\lambda \ell^{\mu} k^\nu.
\end{equation}
is orthogonal to $l^\mu,k^\mu,\eta^\mu$ and it has unit norm and hence is the
other member of the tetrad. Then we have
\begin{equation}
\label{eqtc:7}
J=\frac{1}{16\pi}\int_{\cal S} \eta^{-1/2} \xi^\mu \omega_\mu \, ds.
\end{equation}
We emphasize that this expression is valid only for axially symmetric surfaces.
If we assume vacuum, then we can use the twist potential defined by
(\ref{eq:31}) to obtain
\begin{equation}
\label{eqt:7}
J=\frac{1}{8}\int_{0}^\pi \partial_\theta\omega \, d\theta
=\frac{1}{8} \left(\omega(\pi)-\omega(0) \right).
\end{equation}
Equations (\ref{eqtc:7}) and (\ref{eqt:7}) are the analog to equations
(\ref{eq:37}) and (\ref{eq:36}) for the charge.
We consider now the non-vacuum case. We define the following vector (a part of
the extrinsic curvature of $\mathcal{S}$ with the role of a connection on its normal
cotangent bundle, see e.g. the discussion in \cite{Gou05} \cite{Booth:2006bn})
\begin{equation}
\label{eqt:10}
\Omega^{(\ell)}_\mu = -k^\gamma {q^\lambda}_\mu \nabla_\lambda \ell_\gamma \ .
\end{equation}
Since $\mathcal{S}$ is axially symmetric, the tetrad can be chosen such that
\begin{equation}
\label{eqt:11}
\pounds_\eta l^\mu=\pounds_\eta k^\mu=0.
\end{equation}
In particular this implies that
\begin{equation}
\label{eqt:12}
l^\mu\nabla_\mu\eta^\nu- \eta^\mu\nabla_\mu l^\nu=0.
\end{equation}
Using this equation we obtain that
\begin{equation}
\label{eqt:13}
\eta^\mu \Omega^{(\ell)}_\mu = -k^\mu \eta^\nu \nabla_\nu \ell_\mu=-k^\mu
\ell^\nu \nabla_\nu \eta_\mu.
\end{equation}
From this equation we get the following equivalent expression for the Komar angular
momentum (see, for example, \cite{jaramillo11c})
\begin{equation}
\label{eqt:14}
J=\frac{1}{8\pi}\int_{\cal S} \eta^\mu \Omega^{(\ell)}_\mu \, ds.
\end{equation}
Remarkably, as we will see in the following, the vector $\Omega^{(\ell)}_\mu$
determines even in the non-vacuum case a potential on the surface (we will
denote it by $\bar \omega$) which coincide with the twist potential $\omega$
defined above in the vacuum case. We emphasize that the potential $\bar\omega$
will be only defined on the two-surface $\mathcal{S}$, in contrast to the twist
potential $\omega$ which (in the vacuum case) is defined by equations
(\ref{eqt:3}) and (\ref{eq:31}) on a region of the spacetime.
By construction, the vector $\Omega^{(\ell)}_\mu$ is tangent to the surface
$\mathcal{S}$. Since we have assumed that $\mathcal{S}$ has the $\mathbb{S}^2$ topology, there
exists functions $\hat \omega$ and $\lambda$ such that the vector has the
following decomposition on $\mathcal{S}$ in terms of a divergence-free and an exact form
\begin{equation}
\label{eq:73}
\Omega^{(\ell)}_A=T_A+\nabla_A\lambda, \quad T_A=\epsilon_{AB}\nabla^B\hat \omega.
\end{equation}
The functions $\hat \omega$ and $\lambda$ are fixed up to a constant. In
equation (\ref{eq:73}) we have used the capital indices (which run from $1$ to
$2$) to emphasizes that this is an intrinsic equation on the two-surface $\mathcal{S}$.
By assumption both $\Omega^{(\ell)}_A$ and the intrinsic metric on $\mathcal{S}$ are
axially symmetric, then it follows that the functions $\hat \omega$ and
$\lambda$ are also axially symmetric (i.e. they depend only on $\theta$). In
particular it follows that
\begin{equation}
\label{eq:76}
\xi^AT_A=0, \quad \eta^A\nabla_A \lambda=0,
\end{equation}
where $\xi^A$ is the vector tangent to $\mathcal{S}$ previously defined. Since the norm
$\eta$ is also an axially symmetric function, equation (\ref{eq:76}) implies that
\begin{equation}
\label{eq:77}
T^A D_A\eta =0.
\end{equation}
Equation (\ref{eq:77}) ensures the integrability conditions for the existence
of a function $\bar\omega$ such that
\begin{equation}
\label{eq:78}
T_A=\frac{1}{2\eta}\epsilon_{AB}\nabla^B\bar \omega.
\end{equation}
Note that equation (\ref{eq:78}) is valid only in axial symmetry. Collecting
these results we obtain the decomposition
\begin{equation}
\label{eqt:15}
\Omega^{(\ell)}_A= \frac{1}{2\eta} \epsilon_{A B}\nabla^B \bar \omega
+\nabla_A \lambda.
\end{equation}
In particular, using (\ref{eq:76}), we obtain
\begin{equation}
\label{eq:79}
\eta^A \Omega^{(\ell)}_A= \frac{1}{2\eta} \epsilon_{A B}\eta^A \nabla^B \bar \omega.
\end{equation}
We write equation (\ref{eq:79}) with spacetime indices and we use the
following representation for the tetrad vector $\xi^\mu$
\begin{equation}
\label{eqt:17}
\xi_\nu=\eta^{-1/2} \epsilon_{\mu\nu }\eta^\mu,
\end{equation}
to finally obtain
\begin{equation}
\label{eqt:18}
\eta^\mu \Omega^{(\ell)}_\mu = \frac{1}{2}\eta^{-1/2}\xi^\mu \nabla_\mu\bar \omega.
\end{equation}
Using equation (\ref{eqt:14}) an integrating we finally get
\begin{equation}
\label{eqt:21}
J=\frac{1}{8}\int_{0}^\pi \partial_\theta\bar \omega \, d\theta
=\frac{1}{8} \left(\bar \omega(\pi)-\bar \omega(0) \right).
\end{equation}
The relevance of this construction is that the function $\bar \omega$, which is
only defined at the surface $\mathcal{S}$, plays the role of a potential that can be
defined in the non-vacuum case \cite{Jaramillo:2011pg}. To see the relation between $\bar \omega$ and
$\omega$, note that we have
\begin{equation}
\label{eqt:19}
\xi^\mu\omega_\mu=\xi^\mu \nabla_\mu \bar \omega.
\end{equation}
This equation is valid always (i.e. non-vacuum) and it gives the function
$\bar\omega$ in terms of the twist vector $\omega_\mu$. In the non-vacuum case the
twist potential $\omega$ is not defined but the function $\bar \omega $ is always well
defined. In fact, we can take (\ref{eqt:19}) as definition of $\bar \omega$,
since the right hand side is only a derivative with respect to $\theta$.
On the other hand, in the vacuum case equation (\ref{eqt:19}) implies that
$\omega$ and $\bar \omega$ differs by a constant which is irrelevant since it
does not contribute to the angular momentum.
\section{Mass in axial symmetry}
\label{sec:mass-axial-symmetry}
The main goal of this section is to present the mass formula for axially
symmetric data (\ref{eqq:103}) and the mass functional for two-surfaces
(\ref{eqq:44}). We also discuss their geometrical properties in connection with
harmonic maps.
\subsection{Axially symmetric initial data}
\label{sec:axially-symm-init}
The global geometrical inequality (\ref{eq:42}) is studied on axially
symmetric, asymptotically flat initial data set. In order to present the mass
formula, we first review the basic definitions and properties of this kind of
initial data. For simplicity we concentrate only in the pure vacuum case (see
\cite{Chrusciel:2009ki} \cite{Costa:2009hn} for the electro-vacuum case).
An \emph{initial data set} for the Einstein vacuum equations is given
by a triplet $(S, h_{ij}, K_{ij})$ where $S$ is a connected
three-dimensional manifold, $ h_{ij} $ a Riemannian metric, and $ K_{ij}$ a
symmetric tensor field on $S$.
The fields are assumed to satisfy the vacuum constraint equations
\begin{align}
\label{const1}
D_j K^{ij} - D^i K= 0,\\
\label{const2}
R - K_{ij} K^{ij}+ K^2=0,
\end{align}
where $ {D}$ and $ R$ are the Levi-Civita connection and
the scalar curvature associated with $ {h}_{ij}$, and $ K = K_{ij} h^{ij}$. In
these equations the indices are moved with the metric $ h_{ij}$ and its inverse
$ h^{ij}$.
The initial data is called \emph{maximal} if
\begin{equation}
\label{eq:48}
K=0.
\end{equation}
This condition is crucial because it implies via the Hamiltonian constrain
(\ref{const2}) that the scalar curvature $R$ is non-negative. In fact on the
right hand side of equation (\ref{const2}) it is possible to add a non-negative
matter density and all the arguments behind the proof of theorem \ref{t:main-1}
will also apply since they rely on lower bounds for $R$.
The manifold $S$ is called \emph{Euclidean at infinity}, if there exists a
compact subset $\mathcal{K}$ of $S$ such that $S\setminus \mathcal{K}$ is the
disjoint union of a finite number of open sets $U_k$, and each $U_k$ is
diffeomorphic to the exterior of a ball in $\mathbb{R}^3$. Each open set $U_k$ is called an
\emph{end} of $S$. Consider one end $U$ and the canonical coordinates $x^i$ in
$\mathbb{R}^3$ which contains the exterior of the ball to which $U$ is diffeomorphic
to. Set $r=\left( \sum (x^i)^2 \right)^{1/2}$. An initial data set is called
\emph{asymptotically flat} if $S$ is Euclidean at infinity, the metric $
h_{ij}$ tends to the euclidean metric and $ K_{ij}$ tends to zero as $r\to
\infty$ in an appropriate way. These fall off conditions (see \cite{Bartnik86}
\cite{chrusciel86} for the optimal fall off rates) imply the existence of the
total mass $m$ (or ADM mass \cite{Arnowitt62}) defined at each end $U$ by
\begin{equation}
\label{eq:30adm}
m=\frac{1}{16\pi}\lim_{r\to \infty} \int_{\mathcal{S}_r} \left(\partial_j
h_{ij}-\partial_i
h_{jj}\right ) s^i ds,
\end{equation}
where $\partial$ denotes partial derivatives with respect to $x^i$,
$\mathcal{S}_r$ is the euclidean sphere $r=constant$ in $U$, $s^i$ is its
exterior unit normal and $ds$ is the surface element with respect to the
euclidean metric.
The angular momentum is also defined as a surface integral at infinity
(supplementary fall off conditions must be imposed, see, for example, the
review articles \cite{Szabados04}, \cite{jaramillo11c} and reference
therein). Let $\beta^i$ be an infinitesimal generator for rotations with
respect to the flat metric associated with the end $U$, then the angular
momentum $J$ in the direction of $\beta^i$ is given by
\begin{equation}
\label{eq:35}
J=\frac{1}{8\pi} \lim_{r\to \infty} \int_{\mathcal{S}_r} (K_{ij}
-Kh_{ij})\beta^i s^j ds.
\end{equation}
The above discussion applies to general asymptotically flat initial data. We
have seen that at any end the total mass $m$ and the total angular momentum $J$
are well defined quantities. We consider now axially symmetric initial data.
In analog way as the spacetime definition \ref{d:ax}, we say that the
Riemannian manifold $(S,h_{ij})$ is axially symmetric if its group of
isometries has a subgroup isomorphic to $SO(2)$. We will denote by $\eta^i$
the Killing field generator of the axial symmetry and by $\Gamma$ the axis. The
initial data set is called axially symmetric if in addition $\eta^i$ is also a
symmetry of $K_{ij}$, namely
\begin{equation}
\label{eq:33}
\pounds_\eta K_{ij} =0.
\end{equation}
On an axially symmetric spacetime there exist axially symmetric initial
condition, conversely axially symmetric initial data evolve into an axially
symmetric spacetime. However, on an axially symmetric spacetime it is also
possible to take initial conditions which are not axially symmetric in the
sense defined above. These conditions will of course also evolve into an
axially symmetric spacetime, but the Killing vector is `hidden' on them (these
kind of initial data were studied in \cite{beig97}). We will not consider this
kind of data here since on an axially symmetric spacetime it is always possible
to chose initial conditions which are explicitly axially symmetric in the sense
defined above.
For axially symmetric data we have the Komar integral discussed in
section \ref{sec:angul-moment-axial}. This integral can be calculated in terms
on the initial data if we chose $\mathcal{S}\subset S$.
There exist a very simple relation expression for the Komar integral
on an initial data, namely
\begin{equation}
\label{eq:38}
J=\frac{1}{8\pi}\int_\mathcal{S} K_{ij} \eta^i s^j ds.
\end{equation}
The equivalence between (\ref{eq:38}) and the original definition
(\ref{eqcm:1}) can be seen as follows. Consider the tetrad adapted to $\mathcal{S}$
defined in section \ref{s:potentials}. Assume that $\mathcal{S}\subset S$. By
assumption, the axial Killing vector is tangent to the three-dimensional
surface $S$. Denote by $n^\mu$ the timelike unit normal to the three-surface
$S$, and let $s^\mu$ be the spacelike unit normal to the two-surface $\mathcal{S}$
which lies on $S$. These vectors can be written in term of $\ell^\mu$ and
$k^\mu$ as follows
\begin{equation}
\label{eq:80}
n^\mu= \frac{1}{\sqrt{2}} (\ell^\mu+k^\mu ),\quad s^\mu= \frac{1}{\sqrt{2}} (\ell^\mu-k^\mu ),
\end{equation}
Using these expression, we can write the integrand in
(\ref{eqcm:1}) as
\begin{equation}
\label{eq:81}
J(\mathcal{S})=\frac{1}{8\pi}\int_\mathcal{S} n^\lambda s^\gamma \nabla_\lambda \eta_\gamma.
\end{equation}
The second fundamental form can be written (as spacetime tensor) in terms of
the unit normal $n^\mu$ as
\begin{equation}
\label{eq:82}
K_{\mu\nu}=-h^\lambda_\mu \nabla_\lambda n_\nu.
\end{equation}
Using that $\eta^\mu n_\mu=0$ and that $s^\mu$ is tangent to $S$, from
(\ref{eq:82}) we obtain
\begin{equation}
\label{eq:84}
K_{\mu\nu}s^\mu \eta^\nu= n^\nu s^\lambda \nabla_\lambda \eta_\nu.
\end{equation}
From (\ref{eq:84}) and (\ref{eq:81}) we obtain (\ref{eq:38}).
Comparing expression (\ref{eq:38}) with (\ref{eq:35}) we see that if we chose
in (\ref{eq:35}) the vector $\beta^i=\eta^i$ (that it, the generator of axial
rotations) then these to expression are equivalent since $\eta^is_i=0$ for an
sphere at infinity.
It is possible to calculate the potential $\omega$ for the spacetime Killing
field defined in section \ref{sec:angul-moment-axial} in terms of $K_{ij}$ as
follows. Define the vector $K_i$ by
\begin{equation}
\label{eq:40}
K_i=\epsilon_{ijk}S^j \eta^k, \quad S_i=K_{ij}\eta^j,
\end{equation}
where $\epsilon_{ijk}$ is the volume element with respect to the metric
$h_{ij}$.
Then, as a consequence of equations (\ref{const1}) and (\ref{eq:33}) we have that
\begin{equation}
\label{eq:41}
D_{[i} K_{j]}=0.
\end{equation}
Then the potential $\omega$ is defined by
\begin{equation}
\label{eq:44}
K_i=-\frac{1}{2} D_i \omega.
\end{equation}
It can be checked that this is the same potential as defined in the previous
section (see \cite{Dain:2008xr}) evaluated on the initial surface. The
importance of the expression (\ref{eq:44}) is that allow to calculate $\omega$
in terms of the initial conditions.
The functions $\omega$ and $\eta$ have the information of the stationary part
of the initial conditions. The dynamical part is characterized by the functions
$\eta'$ and $\omega'$ defined by
\begin{equation}
\label{eq:45}
\eta'= -2K_{ij}\eta^i\eta^j, \quad \omega'=\epsilon_{ijk}\eta^i D^j \eta^k.
\end{equation}
The notation comes from the fact that they are related with the time derivative
of this functions, namely if $n^\mu$ is the timelike normal of the initial
data, we have (see \cite{Dain:2008xr})
\begin{equation}
\label{eq:46}
\eta'= n^\mu\nabla_\mu \eta, \quad \omega'= n^\mu\nabla_\mu \omega.
\end{equation}
The whole initial data can be constructed in terms of $(\eta, \omega; \eta',
\omega')$. These functions are the initial data for the wave map equations that
characterize the evolution in axial symmetry. This is clearly seen in the
quotient representation (see \cite{dain10}). For our present purpose the
relevant part of the initial conditions is contained in $(\eta,\omega)$.
So far we have discussed local implications of axial symmetry. Suppose that $S$
is simply connected and asymptotically flat (with, possible, multiple ends).
It can be proved (see \cite{Chrusciel:2007dd}) that in such case the analysis
essentially reduces to consider manifold of the form $S=\mathbb{R}^3\setminus
\sum_{k=0}^{N} i_k $ where $i_k$ are points in $\mathbb{R}^3$. These points represent the
extra asymptotic ends of $S$.
Moreover, in \cite{Chrusciel:2007dd} it has been proved that on $S$ there exists the
following global coordinate systems which will be essential in what follows.
\begin{lemma}[Isothermal coordinates]
\label{isothermal}
Consider an axially symmetric, asymptotically flat (with possible multiple
ends) Riemannian manifold $(S, h_{ij})$, where $S$ is assumed to be simply
connected. Then, there exist a global coordinate system system
$(\rho,z,\varphi)$ such the metric has the following form
\begin{equation}
\label{eq:44b}
h= e^{(\sigma+2q)}(d\rho^2+dz^2)+ \rho^2 e^\sigma (d\varphi + A_\rho
d\rho+ A_z dz)^2,
\end{equation}
where the functions $\sigma,q, A_\rho, A_z$ do not depend on $\varphi$. In
these coordinates, the axial Killing vector is given by $\eta=
\partial/\partial \varphi$ and the square of its norm is given by
\begin{equation}
\label{eq:83}
\eta=\rho^2e^{\sigma}.
\end{equation}
\end{lemma}
Using this coordinates system, the end points $i_k$ are located at the axis
$\rho=0$ of $\mathbb{R}^3$. Define the intervals $I_k$, $1\leq k\leq N-1 $, to be the
open sets in the axis between $i_k$ and $i_{k-1}$, we also define $I_0$ and
$I_N$ as $z< i_0$ and $z >i_N$ respectively (see figure \ref{fig:5}). The
manifold $S$ is Euclidean at infinity with $N+1$ ends. In effect, for each
$i_k$ take a small open ball $B_k$ of radius $r_{(k)}$, centered at $i_k$,
where $r_{(k)}$ is small enough such that $B_k$ does not contain any other
$i_{k'}$ with $k'\neq k$. Let $\bar B_R$ be a closed ball, with large radius
$R$, such that $\bar B_R$ contains all points $i_k$ in its interior. The
compact set $\mathcal{K}$ is given by $\mathcal{K}= \bar B_R \setminus
\sum_{k=1}^N B_k$ and the open sets $U_k$ are given by $B_k\setminus i_k$, for
$1 \leq k \leq N$, and $U_0$ is given by $\mathbb{R}^3 \setminus \bar B_R$. Our choice of
coordinate makes an artificial distinction between the end $U_0$ (which
represent $r\rightarrow \infty$) and the other ones. This is convenient for
our purpose because we want to work always at one fixed end.
We emphasize that the isothermal coordinates $(\rho,z,\varphi)$ (and hence the
corresponding spherical radius $r=\sqrt{\rho^2+z^2}$) are globally defined on
$S$. In what follows we will always use these coordinates.
The smoothness of the initial data on the axis implies that the potential
$\omega$ is constant on each interval $I_k$. These constants are directly
related with the angular momentum of the end points $i_k$. In effect,
the angular momentum of an end $i_k$ is defined to be the Komar integral with
respect to a surface $\mathcal{S}_k$ that encloses only that point. Using the formula
(\ref{eqt:7}) we obtain
\begin{equation}
\label{eq:45gg}
J_k\equiv J(\mathcal{S}_k)=\frac{1}{8}\left (\omega|_{I_k} -\omega|_{I_{k-1}}\right ).
\end{equation}
Let $\mathcal{S}_0$ a surface that enclose all the end points $i_k$, then the total
angular momentum of the end $r\to \infty$ is given by
\begin{equation}
\label{eq:46gg}
J_0\equiv J(\mathcal{S}_0)=\frac{1}{8}\left (\omega|_{I_0} -\omega|_{I_{N}}\right ),
\end{equation}
which is equivalent to
\begin{equation}
\label{eq:47gg}
J_0=\sum_{k=1}^N J_k.
\end{equation}
\begin{figure}
\centering
\includegraphics[width=5cm]{axis-N-extremes}
\caption{Axially symmetric data with $N$ asymptotic ends.}
\label{fig:5}
\end{figure}
The mass (\ref{eq:30adm}) is defined as a boundary integral at infinity. For
axially symmetric initial data there is an equivalent representation of the
mass as a positive definite integral over the three-dimensional slice. This
formula was discovered by Brill \cite{Brill59} and allow him to prove the first
version of the positive mass for pure vacuum axially symmetric gravitational
waves. This formula was generalized to include metrics with non hypersurface
orthogonal Killing vectors (i.e. with non-zero $A_\rho$ and $A_z$ in the
representation (\ref{eq:44b})) in \cite{Gibbons06} and for non-trivial
topologies in \cite{Dain05c}. In particular, this includes the topology of the
Kerr initial data. See also \cite{Dain:2008xr} and \cite{dain10} for related
discussion on this mass formula.
Take isothermal coordinates and assume that the initial data are maximal
(i.e. condition (\ref{eq:48}) holds). Then the total mass (\ref{eq:30adm}) is
given by the following positive definite integral
\begin{equation}
\label{eqq:15}
m= \frac{1}{32\pi}\int_{\mathbb{R}^3} \left[|\partial \sigma |^2
+ \frac{e^{(\sigma+2q)} \omega'^2}{\eta^2} + 2e^{(\sigma+2q)} K^{ij}K_{ij}
\right ] \, dv,
\end{equation}
where $|\partial \sigma |^2= \sigma^2_{,\rho} +\sigma^2_{,z}$ and $dv$ is the
flat volume element in $\mathbb{R}^3$, namely $dv=\rho dz d\rho d\varphi$. Since the
functions are axially symmetric the integration in $\varphi$ is trivial, we
keep however this notation because it will be important in section
\ref{sec:harmonic-maps}.
The remarkable fact of the integral (\ref{eqq:15}) is that it applies also to
the case of multiples ends, the integral is taking over all the extra ends
$i_k$ since it cover the whole $\mathbb{R}^3$.
The integrand in (\ref{eqq:15}) can be further decomposed as follows (see
\cite{dain10} for details)
\begin{equation}
\label{eqq:103}
m= \frac{1}{32\pi}\int_{\mathbb{R}^3} \left[ \frac{e^{(\sigma+2q)}}{\eta^2}\left(\eta'^2 +
\omega'^2 \right)+ F^2 + |\partial
\sigma|^2 + \frac{|\partial \omega|^2}{\eta^2} \right] \, dv .
\end{equation}
Where the term $F^2$ is essentially the square of the second fundamental form of
the two-surface that is obtained by the quotient of the manifold $S$ by the
symmetry $SO(2)$ (this two-surface is a half plane $\rho\geq 0$ minus the end
points $i_k$). The important point in this representation is that it clearly
separates the dynamical and the stationary terms in the mass formula. In fact,
this formula can be seen as the geometrical conserved energy of the wave map
that appear in the evolution equations (see \cite{dain10}).
The first three terms in the integrand of (\ref{eqq:103}) correspond to the
dynamical part for the mass. They vanished for stationary solutions. The last
two terms correspond to the stationary part. They give the total mass for Kerr.
Explicitly the stationary part is given by
\begin{equation}
\label{eqq:5c}
\mathcal{M}(\sigma,\omega)= \frac{1}{32\pi}\int_{\mathbb{R}^3}
\left(|\partial \sigma |^2 + \rho^{-4}e^{-2\sigma} |\partial \omega |^2
\right) dv,
\end{equation}
An hence we obtain the important bound
\begin{equation}
\label{eq:47}
m\geq \mathcal{M}(\sigma,\omega).
\end{equation}
We have presented the mass formula but we have not discussed the asymptotic
conditions that ensures that this integral is well defined. We discuss them now
trying to focus only in the essential aspects, for the technical details we
refer to \cite{Dain06c} and \cite{Chrusciel:2007dd}. All the important
features can be seen already in the term $|\partial \sigma|^2$ in
(\ref{eqq:5c}). We will concentrate in this term only in what follows.
It is instructive to see how the function $\sigma$ behaves for the
Kerr black hole initial data with mass $m$. These
initial data have two asymptotic ends, namely we have only one extra end point
$i_1$. In the limit $r\to \infty$ (that is, the end $U_0$) we have
\begin{equation}
\label{eq:70}
\sigma = \frac{2m}{r}+O(r^{-2}).
\end{equation}
At the end $i_1$ we have
\begin{equation}
\label{eq:74}
\sigma = -4\log(r)+O(1)
\end{equation}
for the non-extreme case, and
\begin{equation}
\label{eq:75}
\sigma = -2\log(r)+O(1),
\end{equation}
for the extreme case. We see that the fall off behavior at the end $i_1$ is
different for the non-extreme and extreme case, this reflect of course the
difference between an asymptotically flat end and a cylindrical end (see
figures \ref{fig:4} and \ref{fig:3} respectively).
To include this kind of fall off behavior let us consider the following
conditions. In the limit $r\to \infty$ we impose
\begin{equation}
\label{eq:69}
\sigma=o(r^{-1/2}), \quad \partial \sigma=o(r^{-3/2}).
\end{equation}
At the end point $i_k$ we impose
\begin{equation}
\label{eq:71}
\sigma =o(r_{(k)}^{-1/2}), \quad \partial \sigma=o(r_{(k)}^{-3/2}),
\end{equation}
Despite the formal similarity of (\ref{eq:71}) and (\ref{eq:69}) they are of a
very different nature. Conditions (\ref{eq:69}) are essentially the standard
asymptotic flat conditions on the metric, they include in particular the
behavior (\ref{eq:70}) of the Kerr initial data. On the other hand, conditions
(\ref{eq:71}) on the extra ends $i_k$ are much more general than asymptotic
flatness. Note that asymptotic flatness at the ends $i_k$ in these coordinates
essentially implies a behavior like (\ref{eq:69}). But conditions (\ref{eq:74})
includes also (\ref{eq:75}) (in fact also any logarithmic behavior with any
coefficient).
The motivation for the fall off conditions (\ref{eq:71}) and (\ref{eq:69}) is
that they ensures that the integral of $|\partial \sigma|^2$ is finite both at
infinity and at the end points $i_k$. And the advantage of condition
(\ref{eq:71}) is that it encompasses both kind of behavior: asymptotically flat and
cylindrical. This is the main motivation of the Brill data definition that
appears in theorem \ref{t2} (see \cite{Dain06c} for details).
\subsection{Harmonic maps}
\label{sec:harmonic-maps}
The crucial property of the mass functional defined in \eqref{eqq:5c}
is its relation to the energy of harmonic maps from $\mathbb{R}^3$ to the
hyperbolic plane $\mathbb{H}^2$: they differ by a boundary term. To see this
relation consider first the mass functional $\mathcal{M}$ defined on a bounded region
$\Omega$ such that $\Omega$ does not intersect the axis $\Gamma$ (given by
$\rho=0$)
\begin{equation}
\label{eqq:3}
\mathcal{M}_{\Omega}= \frac{1}{32\pi}\int_{\Omega}
\left(|\partial \sigma |^2 + \rho^{-4}e^{-2\sigma} |\partial \omega |^2
\right) dv.
\end{equation}
We consider this integral for general functions $\sigma$ and $\omega$ which are
not necessarily axially symmetric.
The function $\log \rho$ is harmonic outside the axis, that is
\begin{equation}
\label{eq:72}
\Delta \log \rho=0, \text{ in } \mathbb{R}^3\setminus\Gamma,
\end{equation}
where $\Delta$ is the flat Laplacian in $\mathbb{R}^3$. Using equation (\ref{eq:72}) we
obtain the following identity
\begin{equation}
\label{eqq:18}
\mathcal{M}_{\Omega}= \mathcal{M}'_{\Omega}- \oint_{\partial \Omega} 4\frac{\partial
\log\rho}{\partial s} (\log\rho+\sigma)\, ds,
\end{equation}
where the derivative is taken in the normal direction to the boundary $\partial
\Omega$ and $\mathcal{M}'_{\Omega}$ is given by
\begin{equation}
\label{eqq:19}
\mathcal{M}'_{\Omega}=\frac{1}{32\pi}\int_{\Omega}
\left(\frac{ |\partial \eta |^2 + |\partial \omega |^2}{\eta^2}
\right) dv.
\end{equation}
Recall that $\eta$ is defined by (\ref{eq:83}).
The functional $\mathcal{M}'_{\Omega}$ defines an energy for maps
$(\eta,\omega):\mathbb{R}^3 \to \mathbb{H}^2$ where $\mathbb{H}^2$ denotes the
hyperbolic plane $\{(\eta,\omega): \eta>0\}$, equipped with the negative constant
curvature metric
\begin{equation}
\label{eqq:53}
ds^2= \frac{d\eta^2+d\omega^2}{\eta^2}.
\end{equation}
The Euler-Lagrange equations for the energy $\mathcal{M}'_{\Omega}$ are given by
\begin{align}
\label{eqq:ha1}
\Delta \log \eta &= -\frac{|\partial \omega|^2}{\eta^2},\\
\label{eqq:ha2}
\Delta \omega & =
2 \frac{\partial \omega\partial \eta}{\eta}.
\end{align}
The solutions of
\eqref{eqq:ha1}--\eqref{eqq:ha2}, i.e, the critical points of
$\mathcal{M}'_{\Omega}$, are called harmonic maps from $\mathbb{R}^3 \to \mathbb{H}^2$.
Since $\mathcal{M}_{\Omega}$ and $\mathcal{M}'_{\Omega}$ differ only by a boundary
term, they have the same Euler-Lagrange equations.
Harmonic maps have been intensively studied, in particular the Dirichlet
problem for target manifolds with negative curvature has been solved
\cite{Hamilton75} \cite{Schoen83} \cite{Schoen82}, and
\cite{Hildebrandt77}. The last article is particularly relevant here. However,
these results do not directly apply in our case because the equations are
singular at the axis. One of the main technical complications in the proof of
theorem \ref{t:main-1} and \ref{t2} is precisely how to handle the singular
behavior at the axis.
We present in what follows the main strategy in the proof of the inequality
(\ref{eq:10}) (we follow the approach presented in \cite{Chrusciel:2007ak} and
\cite{Costa:2009hn}). The core of the proof is the use of a theorem due to
Hildebrandt, Kaul and Widman \cite{Hildebrandt77} for harmonic maps. In that
work it is shown that if the domain for the map is compact, connected, with
nonvoid boundary and the target manifold has negative sectional curvature, then
minimizers of the harmonic energy with Dirichlet boundary conditions exist, are
smooth, and satisfy the associated Euler-Lagrange equations. That is, harmonic
maps are minimizers of the harmonic energy for given Dirichlet boundary
conditions. Also, solutions of the Dirichlet boundary value problem are unique
when the target manifold has negative sectional curvature. Therefore, we want
to use the relation between $\mathcal{M}$ and the harmonic energy $\mathcal{M}'$ in order to
prove that minimizers of $\mathcal{M}'$ are also minimizers of $\mathcal{M}$. There are two
main difficulties in doing this. First, the harmonic energy $\mathcal{M}'$ is not
defined for the functions that we are considering if the domain of integration
includes the axis. Second, we are not dealing with a Dirichlet problem. To
overcome these difficulties an appropriate compact domain is chosen which do
not contain the singularities. Then a partition function is used to
interpolate between extreme Kerr initial data outside this domain and general
data inside, constructing also an auxiliary intermediate region. This solves
the two difficulties in the sense that now the Dirichlet problem on compact
region can be considered, and the harmonic energy is well defined for this
domain of integration. This allows us to show that the mass functional for Kerr
data is less than or equal to the mass functional for the auxiliary
interpolating data. The final step is to show that as we increase the compact
domain to cover all $\mathbb{R}^3$ the auxiliary data converges to the mass functional
for the original general data. This is the subtle and technical part of the proof.
Finally, we mention that it is possible to construct a heat flow for equations
(\ref{eqq:ha1})--(\ref{eqq:ha2}). Note that the first existence result for
harmonic maps used a heat flow \cite{eells64}. In our present setting, an
appropriate heat flow that incorporates the singular boundary conditions is
constructed as follows. Consider functions $(\sigma,\omega)$ which depend on an
extra parameter $t$. Then, we define the following flow
\begin{align}
\label{eq:haf1}
\dot \sigma &= \Delta \sigma +\frac{e^{-2\sigma}|\partial \omega |^2}{\rho ^4},\\
\label{eq:haf2}
\dot \omega & = \Delta \omega -
2 \frac{\partial \omega\partial \eta}{\eta}.
\end{align}
where a dot denotes partial derivative with respect to $t$. Equations
(\ref{eq:haf1})--(\ref{eq:haf2}) represent the gradient flow of the energy
(\ref{eqq:5c}). The important property of the flow is that the energy $\mathcal{M}$ is
monotonic under appropriate boundary conditions. This flow have been used in
\cite{Dain:2009qb} as an efficient method for computing numerically both the
solution and the value of the energy $\mathcal{M}$ at a stationary solution.
\subsection{Two surfaces and the mass functional}
\label{sec:two-surfaces-mass}
In this section we want to define a mass functional over a two-surface
$\mathcal{S}$. This functional plays a major role in the proofs of the quasi-local
inequalities. Let us motivate first the definition.
Consider the mass functional (\ref{eqq:5c}) defined over $\mathbb{R}^3$. Assume that on
$\mathbb{R}^3$ we have a foliation of two-surfaces. For example, take
spherical coordinates $(r,\theta,\varphi)$ and the two-surfaces $r=constant$. We can
split the integral (\ref{eqq:5c}) into an integral over the two surface and a
radial integral. However, the integral over the two-surface alone will not have
any intrinsic meaning since, of course, its integrand depends on $r$. To obtain
an intrinsic expression we need somehow to avoid the $r$ dependence taking some
kind of limit. For extreme Kerr initial data this limit is provided naturally
on the cylindrical end. The functions $\sigma$ and $\omega$ have a well defined
and non-trivial limit there (in contrast with the asymptotic flat end where
they tend to zero). Also, all the radial derivatives go to zero at the
cylindrical end. This surface define an extreme Kerr throat surface (see the
discussion in \cite{Dain:2010uh} and \cite{dain10d} for more details) and it
characterized by only one parameter: the angular momentum
$J$. Explicitly the functions in that limit are given by
\begin{equation}\label{datos}
\sigma_0=\ln(4|J|)-\ln(1+\cos^2\theta),\quad
\omega_0=-\frac{8J\cos\theta}{1+\cos^2\theta}
\end{equation}
And the area of this two-surface is given by
\begin{equation}
A_0=8\pi|J|.
\end{equation}
Consider the following functional over a two-sphere
\begin{equation}
\label{eqq:22}
\mathcal{M}^{\Su}= \int_0^\pi\left( |\partial_\theta \sigma|^2 +4\sigma + \frac{|\partial_\theta
\omega|^2}{\eta^2}\right) \sin\theta \, d\theta.
\end{equation}
We can write this integral as an integral over the unit sphere $\mathbb{S}^2$ in
the following way
\begin{equation}
\label{eqq:44}
\mathcal{M}^{\Su}=\frac{1}{2\pi} \int_{S^2}\left( |D \sigma|^2 +4\sigma + \frac{|D
\omega|^2}{\eta^2}\right) \, ds_0,
\end{equation}
were $ds_0=\sin\theta\, d\theta d\phi$ is the area element of the standard
metric in $\mathbb{S}^2$ and $D$ is the covariant derivative with respect to this
metric. In complete analogy with the discussion in section
\ref{sec:harmonic-maps}, we consider this integral for general functions
$\sigma(\theta, \varphi)$ and $\omega(\theta, \varphi)$ on $\mathbb{S}^2$ which are not
necessarily axially symmetric.
The Euler-Lagrange equations of this functional are given by
\begin{align}
\label{eq:18c}
\Delta_0\sigma-2 & =-\frac{|\partial_\theta \omega|^2}{\tilde\eta^2},\\
\label{eq:20d}
\Delta_0\omega & =2\frac{\partial_\theta\omega\partial_\theta\tilde \eta}{\tilde \eta},
\end{align}
where
\begin{equation}
\label{eq:85}
\tilde \eta=\sin^2\theta \, e^{\sigma}.
\end{equation}
The functional (\ref{eqq:44}) is relevant because the extreme Kerr throat
surface (\ref{datos}) is a solution of the Euler-Lagrange equations. Equations
(\ref{eq:18c})--(\ref{eq:20d}) can be also deduced from the stationary axially
symmetric equations (\ref{eqq:ha1})--(\ref{eqq:ha2}) taking the limit to the
cylindrical end (given by $r\to 0$ in these coordinates) and using the fall off
behavior of the functions at the cylindrical end, namely
\begin{equation}
\label{eq:86}
\sigma =-2\log(r)+ \tilde\sigma(\theta)+O(r^{-1}),\quad \omega=\tilde
\omega(\theta) +O(r^{-1}).
\end{equation}
Note that the $2$ that appears in the second term in the left hand side of
(\ref{eq:18c}) arises from the characteristic $-2\log(r)$ fall off behavior at the
cylindrical end (we have already discuss this property in equation
(\ref{eq:75})). Also this $2$ produces the linear term $4\sigma$ in the mass
functional (\ref{eqq:44}).
The value of the functional (\ref{eqq:44}) at the extreme Kerr throat sphere is
given by
\begin{equation}
\mathcal{M}^{\Su}=8(\ln(2|J|)+1).
\end{equation}
The connection between the mass functional (\ref{eqq:44}) and the energy of
harmonic maps between $\mathbb{S}^2$ and $\mathbb{H}^2$ is very similar as the
one described in the previous section for the mass functional $\mathcal{M}$. Namely,
consider the functional
\begin{equation}
\label{eqq:47}
\mathcal{M}'^{\mathcal{S}}_\Omega= \frac{1}{2\pi} \int_\Omega \frac{|D \eta|^2+|D
\omega|^2}{\eta^2} \, ds_0,
\end{equation}
defined on some domain $\Omega\subset \mathbb{S}^2$, such that $\Omega$ does not include
the poles. Integrating by parts and using the
identity
\begin{equation}
\label{eqq:48}
\Delta_0(\log(\sin\theta))=-1,
\end{equation}
where $\Delta_0$ is the Laplacian on $\mathbb{S}^2$, we obtain the following
relation between $\mathcal{M}$ and $\mathcal{M}'$
\begin{equation}
\label{eqq:49}
\mathcal{M}'^{\mathcal{S}}_\Omega= \mathcal{M}^{\mathcal{S}}_\Omega+4 \int_\Omega \log\sin\theta\, ds_0+
\oint_{\partial
\Omega} (4\sigma + \log\sin\theta) \frac{\partial \log\sin\theta}{\partial
s}\, dl,
\end{equation}
where $s$ denotes the exterior normal to $\Omega$, $dl$ is the line element
on the boundary $\partial \Omega$ and we have used the obvious notation
$\mathcal{M}^{\Su}_\Omega$ to denote the mass functional \eqref{eqq:44} defined over the
domain $\Omega$. The difference between $\mathcal{M}^{\Su}$ and $\mathcal{M}'^{\mathcal{S}}$ are the boundary
integral plus the second term which is just a numerical constant. Note that if
we integrate over $\mathbb{S}^2 $ this constant term is finite
\begin{equation}
\label{eqq:50}
\int_\Omega \log\sin\theta\, ds_0=2\log2-2.
\end{equation}
The boundary terms however diverges at the poles.
In an analogous way as it was described in the previous section, the functional
$\mathcal{M}'^{\mathcal{S}}$ defines an energy for maps $(\eta,\omega): \mathbb{S}^2 \to
\mathbb{H}^2$ where $\mathbb{H}^2$ denotes the hyperbolic plane $\{(\tilde
\eta, \omega ) : \tilde \eta > 0\}$, equipped with the negative constant
curvature metric
\begin{equation}
\label{eqq:51}
ds^2=\frac{d\tilde\eta^2+d\omega^2}{\tilde\eta^2}.
\end{equation}
The Euler-Lagrange equations for the energy $\mathcal{M}'$ are called harmonic maps
from $S^2\to \mathbb{H}^2$. Since $\mathcal{M}^{\Su}$ and $\mathcal{M}'^{\mathcal{S}}$ differ only by a constant
and boundary terms, they have the same Euler-Lagrange equations.
The variational problem for the mass functional on the two-surface is very
similar to the one for the mass functional on $\mathbb{R}^3$ described in section
\ref{sec:harmonic-maps} (see \cite{Acena:2010ws} for the details).
\section{Open Problems}
\label{sec:open-problems}
In this final section I would like to present the main open problems regarding
these geometrical inequalities. In the light of the recent results of
\cite{Hollands:2011sy} there exists now a very interesting open door to higher
dimensions, but this lies out of the scope of the present review and hence in
this section I will restrict myself to four spacetime dimensions. My aim is to
present open problems which are relevant (and probably involve the discovery of
new techniques) and at the same time they appear feasible to solve.
We begin with the global inequality (\ref{eq:42}). The two main open problems
are the following.
\begin{itemize}
\item Remove the maximal condition.
\item Generalization for asymptotic flat manifolds with multiple ends.
\end{itemize}
The situation for the maximal condition in theorems \ref{t:main-1} and \ref{t2}
resembles the strategy of the proof of positive mass theorem by Schoen and Yau
\cite{Schoen79b} \cite{Schoen81} \cite{Schoen81c} \footnote{I thank Marcus
Khuri for pointing this out to me and for relevant discussion on this
subject}. That proof was performed first for maximal initial data and then
extended for general data. It is conceivable that similar techniques (i.e. the
use of Jang equation) can be used here also, but it is far from obvious how to
extend these ideas to the present case.
The most relevant open problem regarding the global inequality (\ref{eq:42}) is
its validity for manifolds with multiple asymptotic ends with non-trivial
angular momentum. The physical heuristic argument presented in section
\ref{sec:physical-picture} applies to that case and hence there little doubt
that the inequality holds. In particular, as we already mentioned in section
\ref{sec:global-inequality}, for the case of three asymptotic ends there are
strong numerical evidences for the validity of the inequality
\cite{Dain:2009qb}.
Theorem \ref{Piotr-Gilbert} (proved in \cite{Chrusciel:2007ak}) reduces the
proof of the inequality to prove the bound (\ref{eq:49}) for the mass
functional evaluated at the stationary solution. The case of three asymptotic
ends (which, roughly speaking, is equivalent to say that we have two black
holes) is special for the following reason. There exists exact stationary
axially symmetric solutions of the Einstein equations (the so-called
double-Kerr-NUT solutions \cite{kramer80b} \cite{neugebauer80}) which represent
two Kerr-like black holes. These solutions contain singularities, that prevent
them to qualify as genuinely equilibrium state for binary black holes. In
fact, one of the main part in the strategy to prove that the uniqueness theorem
hold for the binary case is to prove that these solutions are always singular
\cite{Neugebauer:2011qb}. However, even if these solutions are singular they
can be useful to prove the bound (\ref{eq:49}), because in order to qualify as
a stationary point of the mass functional $\mathcal{M}$ all we need is that the
functions $(\eta, \omega)$ are regular. It is conceivable that some of these
exact solutions have this property (that means, of course, that other
coefficient of the metric are singular) and hence with the explicit expression
for $(\eta, \omega)$ provided by them it will possible to evaluate $\mathcal{M}$ and
check the bound (\ref{eq:49}). In the articles \cite{Manko:2011qh}
\cite{CabreraMunguia:2010uu} the geometrical inequality (\ref{eq:42}) has been
studied for these exact solutions. These results provide a guide for which
subclass of these solutions are potentially useful to prove the bound
(\ref{eq:49}). Unfortunately the solutions, although explicit, are very
complicated and it is very difficult to compute the mass functional $\mathcal{M}$ for
them\footnote{I thank Piotr Chrusciel for relevant discussion on this point}.
To compute the value of $\mathcal{M}$ for this kind of exact solution would be
certainly a very interesting result which not only will prove the inequality
(\ref{eq:42}) for the three asymptotic ends case but also will hopefully
provide some new interpretation of the double-Kerr-NUT solutions. However this
result will be confined to the three asymptotic ends case and probably will not
yield light into the mechanism of the variational problem for the mass
functional $\mathcal{M}$ with multiple ends. The basic property which is expected to
satisfies $\mathcal{M}$ is the following: if an extra black hole is added, with arbitrary
angular momentum, then the value of $\mathcal{M}$ increase. This property is of course
another way of saying that the force between the black holes is always
attractive (in particular, it can not be balanced by spin-spin repulsive
interaction). The variational problem for the mass functional $\mathcal{M}$ with
multiple ends appears to have a remarkably structure. In particular, there is
formal similarity between this problem and the kind of singular boundary
problems for harmonic maps studied in \cite{bethuel94}.
We mention in section \ref{sec:physical-picture} that there is a clear physical
connection between the global inequality (\ref{eq:42}) and the Penrose
inequality with angular momentum in axial symmetry. Hence, it appropriate to
list here the Penrose inequality as a relevant open problem for axially
symmetric black holes (for more detail on this problem see the review
\cite{Mars:2009cj}):
\begin{itemize}
\item Prove the Penrose inequality with angular momentum Eq. (\ref{eq:23}).
\end{itemize}
However, it is important to emphasize that it not clear that the techniques
used to prove theorems \ref{t:main-1} and \ref{t2} will help to solve this
problem. The reason is the following. The proof of the Penrose inequality
involve an inner boundary, namely the black hole horizon. On the other hand,
theorems \ref{t:main-1} and \ref{t2} refer to complete manifolds without inner
boundaries. As we mention in \ref{sec:global-inequality} there are results in
axial symmetry which includes inner boundaries (i.e. \cite{Gibbons06}
\cite{Chrusciel:2011eu}) and use similar techniques as in theorems
\ref{t:main-1} and \ref{t2} (namely, the representation of the mass as positive
definite integral in axial symmetry, see section
\ref{sec:mass-axial-symmetry}). However the boundary for the Penrose inequality
has a very important property: it should be an outer minimal surface (for
simplicity, we discuss only the Riemannian case). This property is very
difficult to incorporate in an standard boundary value problem. To see if the
mass formula in axial symmetry is useful to prove the Penrose inequality with
angular momentum the natural first step is to prove the Riemannian Penrose
inequality without angular momentum in axial symmetry using this mass
formula. The results presented in \cite{Gibbons06} \cite{Chrusciel:2011eu}
contribute in this direction, but so far the problem remains open. May be
there exists a combination of the global flows techniques developed in
\cite{Huisken01} \cite{Bray01} for the Riemannian Penrose inequality with the
mass functional that incorporate the angular momentum in axial symmetry. For
example, it is suggestive that the strategy for the proof of the charged
Penrose inequality Eq. (\ref{eq:22}) given in \cite{Jang79} and
\cite{Huisken01} consists in first prove the inequality (\ref{eq:53}) using a
flow and a lower bound to the scalar curvature that resemble the mass
functional. However, a generalization of this construction to include the
angular momentum is far from obvious.
We turn now to the quasi-local inequalities. The three main problems are the following.
\begin{itemize}
\item Include the charge and the electromagnetic angular momentum in axial
symmetry for the inequality (\ref{e:inequality}).
\item Isoperimetric inequalities in axial symmetry with angular momentum. That
is, a version for theorem \ref{t:main3} (in axial symmetry) with
angular momentum instead of charge.
\item A generalization of the inequality (\ref{e:inequality}) without axial
symmetry.
\end{itemize}
The inclusion of charge in the inequality (\ref{e:inequality}) is important, of
course, since charge is the other relevant parameter that characterized the
Kerr-Newman black hole. But is also relevant for another reason. The angular
momentum that appears in this inequality is the Komar gravitational angular
momentum. In the generalization with charge it is expected that the total
angular momentum (i.e. gravitational plus electromagnetic) appear in this
inequality. This is shown, including also the magnetic charge, in
\cite{gabach11}. In addition, this is important in connection with the
rigidity statement. There is work in progress on this problem \cite{GabJar11}.
We mention in section \ref{sec:physical-picture} that a version of the
inequality (\ref{e:inequality}) for isoperimetric surfaces (instead of black
hole horizons) could have interesting astrophysical applications, since
apparently neutron stars are close to saturate this kind of inequalities. Such
theorem will be analogous to theorem \ref{t:main3} for the charge. However, it
is by no means clear that similar techniques as the one used in the proofs of
theorem \ref{t:main-2} can be applied to that case.
Finally we mention the problem of finding versions of inequality
(\ref{e:inequality}) without any symmetry assumption. In contrast with the
other open problems presented here, this is not a well defined mathematical
problem since there is no unique notion of quasi-local angular momentum in the
general case. However to explore the scope of the inequality in regions close
to axial symmetry (in some appropriate sense) can perhaps provide such a
notion. From the physical point of view I do not see any reason why this
inequality should only hold in axial symmetry.
\section*{Acknowledgments}
I would like to thank Robert Geroch, Jos\'e Luis Jaramillo, Carlos Kozameh and
Walter Simon for illuminating discussions during the preparation of this
review.
The author is supported by CONICET (Argentina). This work was supported in part
by grant PIP 6354/05 of CONICET (Argentina), grant Secyt-UNC (Argentina) and
the Partner Group grant of the Max Planck Institute for Gravitational Physics
(Germany).
|
\section{Introduction}\label{Section_Intro}
This research aims at estimating multivariate functions with the use of the oracle approach. The first step of the method consists in justification of pointwise and global oracle inequalities for the estimation procedure; the second step is the deriving from them adaptive results for estimation of the point functional and the entire function correspondingly. The obtained results show full adaptivity of the proposed estimator as well as its minimax rate optimality.
\paragraph{ Model and set-up}
Let \( \cc D \supset [-1/2, 1/2]^d \) be a bounded interval in \( \RRd \). We observe a path \( \{ Y_{\eps}(t), t \in \cc D \} \), satisfying the stochastic differential equation
\begin{equation}\label{WGN_model}
Y_{\eps}(\dd t) = F(t) \dd t + \eps W(\dd t) \; , \;\; t = (t_1, \ldots, t_d) \in \cc D,
\end{equation}
where \( W \) is a Brownian sheet and \( \eps \in (0,1) \) is the deviation parameter.
\par In the single-index modeling the signal \( F \) has a particular structure:
\begin{equation}\label{single-index}
F(x) = f(x^{\T} \ta^{\circ}),
\end{equation}
where \( f : \RR \to \RR \) is called link function and \( \ta^{\circ} \in \SSd \) is the index vector.
\par
We consider the case of completely unknown parameters \( f \) and \( \ta^{\circ} \) and the only technical assumption is that $f\in\mathbb{F}_M$ where
$\mathbb{F}_M=\left\{g: \RR \to \RR\; |\; \sup_{u\in\RR}|g(u)| \le M\right\}$ for some $M>0$. However, the knowledge of $M$ as well as any information on the smoothness of the link function are not required for the proposed below estimation procedure. The consideration is restricted to the case $d=2$ except the second assertion of Theorem \ref{th:pointwise-adaptation} concerning a lower bound for function estimation at a given point.
Also, without loss of generality we will assume that $\cc D=[-1,1]^{2}$ and $\eps\le e^{-1}$.
\par Let \( \tilde F(\cdot) \) be an estimator, i.e. a measurable function of the observation
\( \{ Y_{\eps}(t), t \in \cc D \} \) and \( \EE_F^{\eps} \) denote the mathematical expectation with respect to \( \PP_F^{\eps} \), the family of probability distributions generated by the observation process \( \{ Y_{\eps}(t), t \in \cc D \} \) on the Banach space of continuous functions on \( \cc D \), when \( F \) is the mean function. The estimation quality is measured by the \( L_r \) risk, \( r \in [1, \infty) \),
\begin{equation}\label{global risk}
\cc R_r^{(\eps)} (\tilde F, F)
=
\expect{F} \| \tilde F - F \|_r,
\end{equation}
where \( \| \cdot \|_r \) is the \( L_r \) norm on \( [-1/2, 1/2]^2 \)
or by the ``pointwise'' risk
\begin{equation}\label{local risk}
\cc R_{r,x}^{(\eps)} (\tilde F, F)
=
(\expect{F} |\tilde F (x) - F(x)|^r )^{1/r}.
\end{equation}
\par The aim is to estimate the entire function \( F \) on
\( [-1/2, 1/2]^2 \) or its value \( F(x) \) from the observation \( \{ Y_{\eps}(t), t \in \cc D \} \) satisfying SDE
\eqref{WGN_model} without any prior knowledge of the nuisance parameters: the function \( f \) and the unit vector \( \ta^{\circ} \). More precisely, we will construct an adaptive (not depending of \( f \) and \( \ta^{\circ} \)) estimator \( \hat F(x) \) at any point \( x \in [-1/2,1/2]^{2} \). In what follows \( \hat F \) notation stands for an adaptive estimator and \( \tilde F \) denotes an arbitrary estimator. Our estimation procedure is a random selector from a special family of kernel estimators parametrized by a window size (bandwidth) \( h>0 \) and a direction of the projection \( \ta \in \SS \), see Section \ref{sect:SelectionRule} below. For that procedure we then establish a pointwise oracle inequality (Theorem \ref{th:local-oracle-inequality}) of the following type:
\begin{equation}\label{eq:intro_pointwise_oracle_ineq_rough}
\cc R_{r,x}^{(\eps)} (\hat F, F)
\le
C_1 \, \eps \sqrt{\ln(1/\eps) /h^{*} (x^{\T} \ta^{\circ})}
+ C_2 \, \eps \sqrt{\ln(1/\eps) },
\end{equation}
where \( h^{*} \) is an optimal in a certain sense (oracle) bandwidths, see Definition~\ref{def_oracle bandwidth}. As
\( r < \infty \) Jensen's inequality and Fubini's theorem trivially imply
\begin{equation*}
\left[ \cc R_r^{(\eps)} (\hat F, F)\right]^r
\le
\expect{F} \big\| \hat F (\cdot) - F(\cdot) \big\|^{r}_r
=
\big\|\cc R_{r,\cdot}^{(\eps)} (\hat F, F)\big\|_r^r.
\end{equation*}
Hence, we immediately obtain the ``global'' oracle inequality
\begin{equation}\label{eq:intro_global_oracle_ineq_rough}
\cc R_{r}^{(\eps)} (\hat F, F)
\le
C_1 \, \eps \| \sqrt{\ln(1/\eps) /h^{*}} \|_r
+ C_2 \, \eps \sqrt{\ln(1/\eps) }.
\end{equation}
Both inequalities \eqref{eq:intro_pointwise_oracle_ineq_rough} and \eqref{eq:intro_global_oracle_ineq_rough} aside of being quite informative itself -- we will see in Section~\ref{OracleEstmator} from Proposition~\ref{prop:risk-of-oracle-estimator} that they claim that our adaptive estimator mimics its ideal (oracle) counterpart, i.e. their risk bounds differ only by a numerical constant, -- they are further used to judge the minimax rate of convergence under the pointwise and \( L_r \) losses correspondingly (Theorems \ref{th:pointwise-adaptation} and \ref{th:global-adaptation}). We will see that these rates are in accordance with Stone's dimensionality reduction principle, see pp. 692-693 in \cite{Stone85}. Indeed, as the statistical model is effectively one-dimensional due to the structural assumption \eqref{single-index} so the rate of convergence is.
\par The obtained results demonstrate full adaptivity of the proposed estimator to the unknown direction of the projection \( \ta^{\circ} \) and the smoothness of \( f \). Moreover, the lower bound given in the second assertion of Theorem~\ref{th:pointwise-adaptation} shows that in the case of pointwise estimation over the range of classes of \( d \)-variate functions having the single-index structure, see definition~\eqref{Def:SI-class_local}, our estimator is even optimally rate adaptive, that is it achieves the minimax rate of convergence. This fact is in striking contrast to the common knowledge that a payment for pointwise adaptation in terms of convergence rate is unavoidable. Indeed, if the index \( \ta^{\circ} \) would be known, than the problem boils down to pointwise adaptation over H\"{o}lder classes in the univariate GWN model. As demonstrated in \cite{Lep1990}, an optimally adaptive estimator does not exist in this case.
\begin{comment}
\paragraph{Objectives}
The goal of our studies is at least threefold. First, we seek an estimation procedure $\hat{F}(x)$ for $F$, which could
be applicable to any function $F$ satisfying (\ref{single-index}) at any given point $ x\in[-1/2,1/2]^{2}$.
Moreover, we would like to bound the quality provided by this estimator uniformly
over the set $\mathbb{F}_M\times\SS$. More precisely, we want to establish for $\hat{F}(x)$ the so-called local oracle inequality: for any
$ x\in[-1/2,1/2]^{2}$
\begin{equation}
\label{eq:local-oracle-intro}
\cc R_{r,x}^{(\eps)} (\hat F, F)\leq C_r A^{(\eps)}_{f,\ta^{\circ}}(x),\quad\forall f\in\mathbb{F}_M,\;\;\forall\ta^{\circ}\in\SS.
\end{equation}
Here the quantity $A^{(\eps)}_{f,\ta^{\circ}}(\cdot)$ is completely determined by the function $f$, vector~$\ta^{\circ}$ and noise level $\eps$,
while $C_r$ is an absolute constant which may depend only on $r$.
Being established the local oracle inequality allows us to derive minimax adaptive results for the function estimation at a given point.
Indeed, let $\mathbb{F}(\gamma),\;\gamma\in\Upsilon,$ be a collection of functional classes such that $\cup_{\gamma\in\Upsilon}\mathbb{F}(\gamma)\subseteq\mathbb{F}_M$. For any $\gamma\in\Upsilon$ define
$$
\phi_\eps(\gamma)=\inf_{\tilde{F}}\sup_{(f,\ta^{\circ})\in\mathbb{F}(\gamma)\times\SS}\cc R_{r,x}^{(\eps)} (\tilde F, F),
$$
where infimum is taken over all possible estimators. The quantity $\phi_\eps(\gamma)$ is the minimax risk on $\mathbb{F}(\gamma)\times\SS$, and
the problem arisen in the framework of minimax adaptive estimation consists in the following. One has to construct an estimator $F^*$ such that
for any $\gamma\in\Upsilon$
\begin{equation}
\label{eq:local-adaptation-intro}
\sup_{(f,\ta^{\circ})\in\mathbb{F}(\gamma)\times\SS}\cc R_{r,x}^{(\eps)} (F^*, F)\;\asymp\; \phi_\eps(\gamma),\;\;\eps\to 0.
\end{equation}
The estimator $F^*$ satisfying (\ref{eq:local-adaptation-intro}) is called adaptive over the class collection $\{\mathbb{F}(\gamma),\;\gamma\in\Upsilon\}$.
Let (\ref{eq:local-oracle-intro}) be proved and suppose additionally that for any $\gamma\in\Upsilon$
$$
\sup_{(f,\ta^{\circ})\in\mathbb{F}(\gamma)\times\SS}A^{(\eps)}_{f,\ta^{\circ}}(x)\asymp\;\phi_\eps(\gamma),\;\;\eps\to 0.
$$
Then one can assert that the estimator $\hat{F}$ is adaptive over the collection $\{\mathbb{F}(\gamma),\;\gamma\in\Upsilon\}$.
Thus, the first task is to prove (\ref{eq:local-oracle-intro}).
Next, we apply this result to minimax adaptive estimation over the collection
$\mathbb{F}(\gamma)=\mathbb{H}(s,L),\; \gamma=(s,L)$, where $\mathbb{H}(s,L)$
is a H\"older class of functions, see Definition \ref{def:holder-class}.
In particular, we find the minimax rate over $\mathbb{H}(s,L)\times \SS $ and
prove that our estimator $\hat{F}$ achieves that rate, i.e. is adaptive. This result is quite surprising because if $\ta^{\circ}$
is fixed, say $\ta^{\circ}=(1,0)^{\T}$, then it is well known that an adaptive estimator does not exist, see \cite{Lep1990}.
\smallskip
Note also that local oracle inequality (\ref{eq:local-oracle-intro}) allows us to bound from above the ``global'' risk as well.
Indeed, in view of Jensen's inequality and Fubini's theorem
$$
\left[ \cc R_r^{(\eps)} (\hat F, F)\right]^r\leq \expect{F} \big\| \hat F (\cdot) - F(\cdot) \big\|^{r}_r=\big\|\cc R_{r,\cdot}^{(\eps)} (\hat F, F)\big\|_r^r
$$
and, therefore,
\begin{equation}
\label{eq:global-oracle-intro}
\cc R_r^{(\eps)} (\hat F, F)\leq C_r \big\|A^{(\eps)}_{f,\ta^{\circ}}(\cdot)\big\|_r.
\end{equation}
The latter inequality is called global oracle inequality. As local oracle inequality (\ref{eq:local-oracle-intro})
is a powerful tool for deriving minimax adaptive results in pointwise estimation, so global oracle inequality
(\ref{eq:global-oracle-intro}) can be used for constructing adaptive estimators of the entire function \( F \).
We will consider the collection of Nikol'skii classes $\mathbb{N}_p(s,L)$, see Definition~\ref{def:nikolskii-class},
where $s,L>0$ and $1\leq p<\infty$. It is important to emphasize that by considering these classes we want to estimate
the functions possessing inhomogeneous smoothness.
This means that the underlying function can be very regular on some parts of the observation domain and rather irregular on the other sets.
We will compute the asymptotic bounds on
$$
\sup_{(f,\ta^{\circ})\in \mathbb{N}_p(s,L)\times\SS}\big\|A^{(\eps)}_{f,\ta^{\circ}}(\cdot)\big\|_r
$$
and show that if $(2s+1)p<r$ the rate of convergence coincides with the minimax rate over $\mathbb{N}_p(s,L)\times\SS$. This means that our estimator $\hat{F}$
is adaptive over the collection $\left\{\mathbb{N}_p(s,L)\times\SS,\; s>0,L>0\right\}$ whenever $(2s+1)p<r$.
In the case $(2s+1)p\geq r$ we will prove that the latter bound
differs from the bound on the minimax risk by logarithmic factor.
Following the contemporary language we say that the estimator $\hat{F}$ is ``nearly'' adaptive.
We note, nevertheless, that the construction of adaptive estimator in the case $(2s+1)p\geq r$
under single-index constrain (\ref{single-index}) remains an open problem.
\paragraph{Remarks} It turns out that the adaptation to the unknown
\( \ta^{\circ} \) and \( f(\cdot) \) can be formulated in terms of selection from a special family of
kernel estimators in the spirit of the Lepski and the Goldenschluger-Lepski selection
rules, see \cite{Lep1990, KLP2001, GoldLep2008}. However the proposed here procedure is quite different from
the aforementioned ones, and it allows us to solve the problem of minimax adaptive estimation under the \( L_r \) losses
over a collection of Nikol'skii classes.
\par It is worth mentioning that the considered single-index model is not only of high theoretical
interest but is also actively used especially in econometrics, e.g. \cite{Horowitz1998, MADDALA}. The estimation, nevertheless, is usually performed under smoothness
assumptions on the link function. One usually uses the \( L_2 \) losses, and the available methodology is based on
these restrictions. To our best knowledge the only exceptions are \cite{Golubev1992} for the minimax estimation under the
projection pursuit constrains, and a recent paper \cite{GoldLep2009} presenting a novel procedure permitting to adapt
simultaneously to unknown smoothness and structure.
\end{comment}
\par Although the literature on the single-index model is rather numerous, we mention only books \cite{Haerdle2004}, \cite{Horowitz1998}, \cite{Gyorfi2002} and \cite{Koros}, quite a few works address the problem of function estimating when both the link function and index are unknown. To the best of our knowledge the only exceptions are \cite{Golubev1992}, \cite{Lecue} and \cite{GoldLep2008}. An adaptive projection estimator is constructed in \cite{Golubev1992}, in \cite{Lecue} the aggregation method is used. Both the papers employ \( L_2 \) losses. \cite{GoldLep2008} seems to be the first work on pointwise adaptive estimation in the considered set-up, the upper bound for estimation at a point obtained therein is similar to our, but the estimation procedure is different.
\paragraph{Organization of the paper} In Section \ref{subsec:oracle-approach} we motivate and explain the proposed
selection rule. Then in Section~\ref{subsec:OI} we establish for it local and global oracle inequalities of type \eqref{eq:intro_pointwise_oracle_ineq_rough} and \eqref{eq:intro_global_oracle_ineq_rough}. In Section \ref{subsec:adaptive-estimation} we apply these results to minimax adaptive estimation. Particularly, Section~\ref{subsec:PointwiseAdaptation} is devoted to the upper bound and already discussed above lower bound for estimation over a range of H\"{o}lder classes. Section \ref{sect:Global_Adapt} addresses the ``global'' adaptation under the \( L_r \) losses and the estimator performance over the collection of classes of single-index functions with the link function in a Nikol'skii class, see Definition~\ref{def:nikolskii-class} and~\eqref{Def:SI-class_global}. That consideration allows to incorporate in analysis functions of inhomogeneous smoothness, that is those which can be very smooth on some parts of observation domain and irregular on the others. The proofs of the main results are given in Section~\ref{sec:proofs} and the proofs of technical lemmas are postponed until Appendix.
\begin{comment}
\section{Main results}
\label{sec:main-results}
Then, we apply these results to adaptive estimation over a collection of H\"older classes (pointwise estimation) and over a collection of
Nikol'skii classes (estimating the entire function with the accuracy of an estimator measured under the~$L_r$~risk).
\end{comment}
\section{Oracle approach}
\label{subsec:oracle-approach}
Below we define an ``ideal'' (oracle) estimator and describe our estimation procedure. Then we present local and global oracle inequalities demonstrating a nearly oracle performance of the proposed estimator.
\par Denote by $\K:\RR \to \RR$ any function (kernel) that integrates to one, and define for any $z\in \RR$, $h\in(0,1]$ and any $f\in\mathbb{F}_M$
$$
\Delta_{\K,f}(h,z)
=
\sup_{\delta\le h}\left| \frac{1}{\delta} \int\K\Big( \frac{u-z}{\delta} \Big)\big[f(u)-f(z)\big]\dd u\right|,
$$
a monotonous approximation error of the kernel smoother
\( 1/\delta\int\K\big[ (u-z) /\delta\big] f(u)\dd u \). In particular, if the function $f$ is uniformly continuous then\( \Delta_{\K,f}(h,z)\to 0 \) as \( h\to 0 \).
In what follows we assume that the kernel $\K$ obeys
\begin{assumption}
\label{ass:assumption-on-kernel}
(1)\; $\text{supp}(\K)\subseteq[-1/2,1/2]$, $\int \K=1$, $\K$ is symmetric;
\medskip
\hskip2.4cm (2)\; there exists $Q>0$ such that
\smallskip
\hskip3.2cm$
\big|\K(u)-\K(v)\big|\leq Q|u-v|,\quad\forall u,v\in \RR.
$
\end{assumption}
\subsection{Oracle estimator}\label{OracleEstmator} For any $y\in \RR$ denote by
$$
\overline{\Delta}_{\K,f}(h,y)=\sup_{a>0} \frac{1}{2a} \int_{y-a}^{y+a}\Delta_{\K,f}(h,z)\dd z,
$$
the Hardy-Littlewood maximal function of $\Delta_{\K,f}(h,\cdot)$, see for instance \cite{WheedenZygmund1977}.
Put also $\Delta^{*}_{\K,f}(h,\cdot)=\max\left\{\overline{\Delta}_{\K,f}(h,\cdot),\Delta_{\K,f}(h,\cdot)\right\}$
and remark that in view of the Lebesgue differentiation theorem
$\Delta^{*}_{\K,f}(h,\cdot)$ and $\overline{\Delta}_{\K,f}(h,\cdot)$ coincide almost everywhere.
Note also, that if $f$ is a continuous function then $\Delta^{*}_{\K,f}(h,\cdot)\equiv \overline{\Delta}_{\K,f}(h,\cdot)$.
Define for $\forall y\in \RR$ the oracle (depending on the underlying function) bandwidth \( h^*_{\K,f}(y) \)
\begin{equation}
\label{def_oracle bandwidth}
h^*_{\K,f}(y)=
\sup\big\{h\in [\eps^2, 1]:\quad \sqrt{h}\;\Delta^{*}_{\K,f}(h,y)\leq
\| \K \|_{\infty}\eps \sqrt{\ln (1/\eps)}\big\}.
\end{equation}
We see that, with the proviso that \( f \in \mathbb{F}_M \), the ``bias'' \( \Delta^{*}_{\K,f}(h,\cdot)\le 2M\|\K\|_1 \), and consequently the set \eqref{def_oracle bandwidth} is not empty for all $\eps\leq \exp{\big\{ -(2M\|\K\|_1\big/\|\K\|_\infty)^2\big\}}$. Here \( \| \K \|_p \), \( 1 \le p \le \infty \), denotes the \( L_p \) norm of \( \K \).
\begin{comment}
Moreover, Assumption \ref{ass:assumption-on-kernel} ({\it 2}) implies that $\Delta^{*}_{\K,f}(\cdot,y)$ is continuous on $[\eps^2, 1]$, hence
\begin{eqnarray}
\label{eq1:def_oracle bandwidth}
&\text{either}&\quad\sqrt{ h^*_{\K,f}(y)}\;\Delta^{*}_{\K,f}\big( h^*_{\K,f}(y),y\big)=\| \K \|_{\infty}\eps \sqrt{\ln (1/\eps)},
\\*[2mm]
\label{eq2:def_oracle bandwidth}
&\text{or}&
\quad \sqrt{h}\;\Delta^{*}_{\K,f}(h,y)\leq
\| \K \|_{\infty}\eps \sqrt{\ln (1/\eps)}, \;\;\forall h\in [\eps^2, 1].
\end{eqnarray}
\end{comment}
\begin{comment}
The quantity, similar to the defined above $h^*_{\K,f}(\cdot)$, first appeared in \cite{LepMamSpok97}
in the context of the estimating univariate functions possessing inhomogeneous smoothness.
Some years later this approach has been developed in \cite{LepskiLevit99}, \cite{KLP2001} and \cite{GoldLep2008}
for the estimation of multivariate function.
In these papers, the interested reader can find a more detailed discussion of the oracle approach. In the present paper we try to adopt the ``ideology'' proposed in the aforementioned papers to the estimation under single index constraint. Our main idea is based on rather simple observation.
\end{comment}
For any $(\theta,h)\in\SS\times[\eps^2, 1]$ define the matrix
$$
E_{(\theta,h)}=\left(
\begin{array}{ll}
h^{-1}\theta_1\quad &h^{-1}\theta_2
\\
-\theta_2\quad &\;\theta_1
\end{array}
\right)
$$
and consider the family of kernel estimators
\begin{equation*}
\label{family of estimators}
\cc F
=
\Big\{ \hat F_{(\ta, h)} (\cdot)
=
\det\big(E_{(\ta, h)}\big)\int K\big(E_{(\ta, h)}(t- \cdot)\big) Y_{\eps}(\dd t),\;\; (\theta,h)\in\SS\times[\eps^2, 1]\Big\}.
\end{equation*}
We use the product type kernels $K(u,v)=\K(u)\K(v)$ with a one-dimensional kernel $\K$ obeying Assumption \ref{ass:assumption-on-kernel}. Note also that
$\det\big(E_{(\ta, h)}\big)=h^{-1}$ and
\begin{equation}
\label{eq:distribution-of-kernel-estimator}
\hat F_{(\ta, h)} (\cdot)-\mathbb{E}^{\eps}_F\left[\hat{F}_{(\ta, h)} (\cdot)\right]\quad\sim\quad \mathcal{N}\left(0,\|\K\|^4_2\eps^{2} h^{-1}\right).
\end{equation}
The choice $\theta=\ta^{\circ}$ and $h=h^*:=h^*_{\K,f}(x^{T}\ta^{\circ})$ leads to the ``ideal'' (oracle) estimator $\hat F_{(\ta^{\circ}, h^*)}$, that is the estimator constructed {\it as if} \( \ta^{\circ} \) and \( f \) would be known. Such an ``estimator'' is not available but serves as a quality benchmark, given by the following result.
\begin{proposition}
\label{prop:risk-of-oracle-estimator}
For any \( (f,\ta^{\circ})\in\mathbb{F}_M\times\SS \), \( \eps\leq \exp{\big\{ -\max [1, (2M\|\K\|_1\big/\|\K\|_\infty)^2]\big\}} \)
and any \( r\ge 1 \)
\begin{equation*}
\cc R_{r,x}^{(\eps)} \big(\hat{F}_{(\ta^{\circ}, h^*)}, F\big)\leq \mathfrak{c}_r
\left[\frac{\| \K \|^{4}_{\infty}\;\eps^{2} \ln (1/\eps)}{h^*_{\K,f}(x^{\T}\ta^{\circ})}\right]^{1/2}, \forall x\in[-1/2,1/2]^{2},
\end{equation*}
\end{proposition}
\noindent where $\mathfrak{c}_r=\left[\mathbb{E}\big(1+|\varsigma|\big)^{r}\right]^{1/r},\; \varsigma\sim \mathcal{N}(0,1)$.
The proof is straightforward and can be omitted.
The meaning of Proposition \ref{prop:risk-of-oracle-estimator} is
that the ``oracle'' knows the exact value of the index $\ta^{\circ}$ and the optimal, up to $\ln(1/\eps)$, bias-variance trade-off $h^*$
between the approximation error caused by $\Delta^{*}_{\K,f}(h^{*},\cdot)$ and the variance, see formula~\eqref{eq:distribution-of-kernel-estimator}, of the kernel estimator from the collection
$\mathcal{F}$
\par Below we will propose an adaptive (not depending of \( \ta^{\circ} \) and \( f \) ) estimator and show that this estimator is as good as the oracle one, i.e. that the risk of that estimator is worse than that of Proposition \ref{prop:risk-of-oracle-estimator} by a numerical constant only.
\begin{comment}
\par In the next paragraph we propose a ``real'' (based on the observation) estimator $\hat{F}(\cdot)$, which mimics the oracle estimator.
This means that for any $(f,\ta^{\circ})\in\mathbb{F}_M\times\SS$, $x\in[-1/2,1/2]^{2}$,
$\eps\leq \exp{\big\{ -\max [1, (2M\|\K\|_1\big/\|\K\|_\infty)^2]\big\}}$, and $r>0$
$$
\cc R_{r,x}^{(\eps)} \big(\hat{F}, F\big)\leq C_r\left[\frac{\| \K \|^{4}_{\infty}
\;\eps^{2} \ln (1/\eps)}{h^*_{\K,f}(x^{\T}\ta^{\circ})}\right]^{1/2},
$$
where $C_r$ is an absolute constant independent of the noise level $\eps$ and the underlying function $F$.
The latter result is a local oracle inequality.
The construction of the estimator $\hat{F}(\cdot)$ is based on the data-driven selection from the family $\mathcal{F}$.
\end{comment}
\subsection{Selection rule}\label{sect:SelectionRule} The procedure below is based on a pairwise comparison of the estimators from \( \cc F \) with an auxiliary estimator defined as follows. For any $\ta,\nu\in\SS$ and any $h\in[\eps^2,1]$ introduce the matrices
\begin{equation*}
\overline{E}_{(\theta,h)(\nu,h)}=\left(
\begin{array}{ll}
\frac{(\theta_1+\nu_1)}{2h(1+|\nu^\T\theta|)}\quad &\frac{(\theta_2+\nu_2)}{2h(1+|\nu^\T\theta|)}
\\*[2mm]
-\frac{(\theta_2+\nu_2)}{2(1+|\nu^\T\ta|)}\quad &\;\frac{(\theta_1+\nu_1)}{2(1+|\nu^\T\ta|)}
\end{array}
\right), \quad
E_{(\theta, h)(\nu,h)}=\left\{
\begin{array}{ll}
\overline{E}_{(\theta, h)(\nu,h)},\;\quad&\nu^\T\ta\geq 0;
\\*[2mm]
\overline{E}_{(-\theta, h)(\nu,h)},\quad&\nu^\T\ta< 0.
\end{array}
\right.
\end{equation*}
It is easy to check that
$
(4h)^{-1}\leq\det\big(E_{(\theta,h)(\nu,h)}\big)\leq (2h)^{-1}.
$
Then, similarly to the construction of the estimators from \( \cc F \) we define a kernel estimator parametrized by \( E_{(\ta, h)(\nu, h)} \)
\begin{equation}\label{pseudo-estimator}
\hat F_{(\ta, h)(\nu, h)}(x)
=
\det \big(E_{(\ta, h)(\nu,h)}\big)
\int K( E_{(\ta, h)(\nu,h)}(t-x))Y_{\eps}(\dd t).
\end{equation}
Put $\Lambda(\K,Q)=8\sqrt{\ln{(1+ 2Q\|\K\|_\infty)}}+50$ and let for any $\eta\in (0,1]$
$$
\TH(\eta)=2\| \K \|^{2}_{\infty}\left[\Lambda(\K,Q)+\sqrt{4r+2}+1\right]\eps\sqrt{\eta^{-1}\ln (1/\eps)}.
$$
Set
$\mathcal{H}_\eps=\big\{h_k=2^{-k},\; k = 0,1, \ldots\big\}\cap[\eps^2,1]$ and define for any \(\ta\in \SS \)
and \(h \in \mathcal{H}_\eps \)
\begin{equation}\label{eq:proced_first-step}
R_{(\ta, h)}(x)
=
\sup_{\eta\in\mathcal{H}_\eps:\; \eta\le h}
\Big\{
\sup_{\nu\in\SS}
\big|\hat F_{(\ta, \eta)(\nu, \eta)}(x) - \hat F_{(\nu, \eta)}(x) \big| -\TH(\eta)
\Big\}.
\end{equation}
\noindent For any $x\in[-1/2,1/2]^{2}$ introduce the random set
\begin{eqnarray*}
\cc P (x)
=
\big\{ (\ta, h)\in\SS\times\mathcal{H}_\eps : R_{(\ta, h)}(x) \le 0 \big\},
\end{eqnarray*}
and let $\tilde{h}=
\max\big\{h:\;\; (\theta,h)\in\cc P (x)\big\}$ if $\cc P (x)\neq\emptyset.$ Note that there exists $\vartheta\in\SS$
such that $(\vartheta,\tilde{h})\in\mathcal{P}(x)$, since the set $\mathcal{H}_\eps$ is finite.
Define
\begin{eqnarray*}
\hat{\ta}=\left\{
\begin{array}{ll}
(1,0)^{\T},\quad & \cc P (x)=\emptyset;
\\*[1mm]
\theta \text{ s.t. } (\theta,\tilde{h})\in\cc P (x),\quad& \cc P (x)\neq\emptyset.
\end{array}
\right.
\end{eqnarray*}
If $\hat{\theta}$ is not unique, let us make any measurable choice. In particular, if $\hat{\Theta}:=\big\{\theta\in\SS:\;(\theta,\tilde{h})\in\cc P (x)\big\}$ one can choose $\hat{\theta}$ as a vector belonging to
$\hat{\Theta}$ with the smallest first coordinate. The measurability of this choice follows from the fact that the mapping $\ta\mapsto R_{(\ta, h)}(x)$ is almost surely continuous on $\SS$. This continuity, in its turn, follows from Assumption \ref{ass:assumption-on-kernel} ({\it 2}), bound (\ref{eq100:proof-lemma-gauss-on-matrices}) for Dudley's entropy integral proved in Lemma \ref{lem:gauss-on-matrices} below and the condition $f\in\mathbb{F}_M$.
Define
\begin{equation}\label{eq:proced_second-step}
\hat{h}=\sup\left\{h\in\mathcal{H}_\eps:\;\;\left|\hat F_{(\hat{\ta},h)}(x)- \hat F_{(\hat{\ta}, \eta)}(x)\right|
\leq\TH(\eta),\;\;\forall \eta\leq h,\;\eta\in\mathcal{H}_\eps\right\}
\end{equation}
and put as a final estimator
\( \hat F (x)=\hat F_{(\hat \ta, \hat h)}(x)\).
\par The proposed above procedure belongs to the stream of pointwise adaptive procedures originating from \cite{Lep1990}. Indeed, the second step determined by \eqref{eq:proced_second-step} for the ``frozen'' \( \hat \ta \) is exactly the procedure of \cite{Lep1990} which was originally developed in the framework of the univariate GWN model. There is a rather vast literature on that topic, we mention \cite{BauerHohageMunk} adapted the method of \cite{Lep1990} for the choice of the parameter for iterated Tikhonov regularization in nonlinear inverse problems, \cite{BertinRivoirard} showed the maxiset optimality of that procedure for bandwidth selection under the \( sup \) norm losses, \cite{Chichignoud} used it for selecting among local bayesian estimators, \cite{Gaiffas} studied the problem of pointwise estimation in random design Gaussian regression, \cite{Serdyukova} investigated a heteroscedastic Gaussian regression under noise misspecification, among many others.
\par The application of \cite{Lep1990} requires some sort of ordering on the set of estimators, for instance in~\eqref{eq:proced_second-step} as soon as \( \hat \ta \) is fixed it is due to the monotonicity of the ``bias'' \( \Delta^{*}_{\K,f}(\cdot,y) \). However, when the projection direction is unknown no natural order on \( \cc F \) is available. This problem is similar to the one arising in generalizations of the pointwise adaptive method for multivariate (anisotropic) settings, see for developments in that direction \cite{LepskiLevit99}, \cite{KLP2001} and \cite{GoldLep2009}. Usually the aforementioned issue requires to introduce an auxiliary estimator and construct a procedure carefully capturing the ``incomparability'' of the estimators. In the considered set-up it is realized by the first step of procedure with \( R_{(\ta, h)}(x) \) given by \eqref{eq:proced_first-step}.
\subsection{Oracle inequalities}\label{subsec:OI}
Throughout the paper we assume that
$$
\eps\leq \exp{\big\{ -\max [1, (2M\|\K\|_1\big/\|\K\|_\infty)^2]\big\}}.
$$
\begin{theorem}
\label{th:local-oracle-inequality}
For any $(f,\ta^{\circ})\in\mathbb{F}_M\times\SS$, $x\in[-1/2,1/2]^{2}$ and any $r\ge 1$
$$
\cc R_{r,x}^{(\eps)} \Big(\hat{F}_{(\hat{\theta}, \hat{h})}, F\Big)
\leq
C_{r,1}(Q,\K)\sqrt{\frac{\|\K\|^4_\infty\eps^{2} \ln (1/\eps)}{h^*_{\K,f}(x^{T}\ta^{\circ})}}
+
C_{r,2}(M,Q,\K)\| \K \|^2_{\infty}\eps\sqrt{\ln (1/\eps)}.
$$
The constants $C_{r,1}(Q,\K)$ and $C_{r,2}(M,Q,\K)$ are given in the beginning of the proof.
\end{theorem}
As already mentioned, the global oracle inequality is
obtained by integrating the local oracle inequality. For ease of notation, we write \( r(\eps) = C_{r,2}(M,Q,\K)\| \K \|^2_{\infty}\,\eps\sqrt{\ln (1/\eps)} \) and $C_r=C_{r,1}(Q,\K)$. It follows from Jensen's inequality and Fubini's theorem that
$$
\cc R_r^{(\eps)} (\hat F, F)
\leq
\big\|\cc R_{r,\cdot}^{(\eps)} (\hat F, F)\big\|_r
\leq
C_r \left\{ \int_{[-1/2,1/2]^{2}}\left[\frac{\|\K\|^4_\infty\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(x^{T}\ta^{\circ})}\right]^{\frac{r}{2}}\dd x \right \}^{\frac{1}{r}} + r(\eps) .
$$
Integration by substitution gives:
$$
\int_{[-1/2,1/2]^{2}}\left[\frac{\|\K\|^4_\infty\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(x^{T}\ta^{\circ})}\right]^{\frac{r}{2}}\dd x\leq \int_{-1/2}^{1/2}\left[\frac{\|\K\|^4_\infty\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(z)}\right]^{\frac{r}{2}}\dd z
$$
leading to the following result.
\begin{theorem}
\label{th:global-oracle-inequality}
For any $(f,\ta^{\circ})\in\mathbb{F}_M\times\SS$ and any $r\ge1$
$$
\cc R_r^{(\eps)} \Big(\hat{F}_{(\hat{\theta}, \hat{h})}, F\Big)\leq C_{r,1}(Q,\K)\left\|\sqrt{\frac{\|\K\|^4_\infty\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(\cdot)}}\right\|_r+C_{r,2}(M,Q,\K)\| \K \|^2_{\infty}\eps\sqrt{\ln (1/\eps)}.
$$
\end{theorem}
\begin{comment}
\subsubsection{Extension to the case $\ta^{\circ}\notin\SS $}
\label{sec:extensions }
Define $f_{\ta^{\circ}}(t)=f\big(|\ta^{\circ}|_2t\big)$, $\vartheta^*=\ta^{\circ}\big/|\ta^{\circ}|_2$ and let $F_{\ta^{\circ}}(x):=f_{\ta^{\circ}}\big(x^\T\vartheta^*\big)$.
Obviously,
$$
f_{\ta^{\circ}}\big(x^\T\vartheta^*\big)=f\big(x^\T\ta^{\circ}\big),\;\forall x\;\;\Rightarrow\;\; F_{\ta^{\circ}}(\cdot)\equiv F(\cdot)
$$
and the estimation of $F(\cdot)$ is equivalent to the estimation of $F_{\ta^{\circ}}(\cdot)$. Moreover $\vartheta\in\SS$ and, therefore,
results obtained in Theorems \ref{th:local-oracle-inequality} and \ref{th:global-oracle-inequality} are applicable. To do that it suffices
to replace $f$ by $f_{\ta^{\circ}}$ in the definition of $h^*_{\K,f}(\cdot)$.
We note, however, that there is no any general receipt of expressing $h^*_{\K,f_{\ta^{\circ}}}(\cdot)$ via $h^*_{\K,f}(\cdot)$, although in particular cases (mainly in adaptive estimation over the collection of classes of smooth functions) it is often possible.
Consider the following example. Let $f$ satisfy the H\"older condition
$
|f(x)-f(y)|\leq L|x-y|^{\alpha},\;\;\forall x,y\in\RR,
$
for some $0<\alpha\leq 1$ and $L>0$. In this case
$$
|f_{\ta^{\circ}}(x)-f_{\ta^{\circ}}(y)|\leq L|\ta^{\circ}|^{\alpha}_2|x-y|^{\alpha}\;\;\forall x,y\in\RR.
$$
For any $h>0$ this yields $\Delta^{*}_{\K,f_{\ta^{\circ}}}(h,\cdot)\leq C(\cc K, \alpha)L|\ta^{\circ}|^{\alpha}_2h^{\alpha}$ with constant \( C(\cc K, \alpha) = \int|\cc K| |t|^\alpha \dd t \) , and, therefore,
$$
h^*_{\K,f_{\ta^{\circ}}}(\cdot)
\geq
\big(|\ta^{\circ}|_2\big)^{-\frac{2\alpha}{2\alpha+1}}\left(\frac{\|\K\|_\infty\eps
\sqrt{\ln (1/\eps)}}{L C(\cc K, \alpha)}\right)^{\frac{2}{2\alpha+1}}.
$$
\end{comment}
\section{Adaptation}
\label{subsec:adaptive-estimation}
In this section with the use of the local oracle inequality from Theorem~\ref{th:local-oracle-inequality} we solve the problem of pointwise adaptive estimation over a collection of H\"older classes. Then, we turn to the problem of adaptive estimating the entire function over a collection of Nikol'skii classes with the accuracy of an estimator measured under the~$L_r$~risk. That is done with the help of the global oracle inequality given in Theorem~\ref{th:global-oracle-inequality}.
\par Throughout this section we will assume that the kernel $\K$ satisfies additionally Assumption \ref{ass2:assumption-on-kernel} below.
Introduce the following notation: for any $a>0$ let $m_a \in\mathbb{N}$ be the maximal integer
strictly less than $a$.
\begin{assumption}
\label{ass2:assumption-on-kernel}
There exists $\bob>0$ such that
$$
\int z^{j}K(z)\dd z=0,\;\;\forall j=1,\ldots,m_{\bob}.
$$
\end{assumption}
\subsection{Pointwise adaptation}\label{subsec:PointwiseAdaptation}
Let us firstly recall the definition of H\"olderian functions.
\begin{definition}
\label{def:holder-class}
Let \( \beta>0 \) and \( L>0 \). A function $g:\mathbb{R}\to\RR$ belongs to the H\"older class $\mathbb{H}(\beta,L)$ if $g$ is $m_\beta$-times continuously
differentiable, $\|g^{(m)}\|_\infty\leq L,\;\forall m\leq m_\beta,$ and
$$
\left|g^{(m_\beta)}(t+h)-g^{(m_\beta)}(t)\right|\le L h^{\beta-m_\beta},\;\;\forall t \in \RR \; \text{and} \; h>0.
$$
\end{definition}
\par The aim is to estimate the function $F(x)$ at a given point $x\in [-1/2,1/2]^2$ under the additional assumption that $F\in\mathbb{F}(\bob):=\bigcup_{\beta\leq\bob}\bigcup_{L>0}\mathbb{F}_2(\beta,L)$, where
\begin{equation}\label{Def:SI-class_local}
\mathbb{F}_d(\beta,L)=\left\{F:\RR^d\to\RR\;|\; F(z)=f\big(z^{\T}\theta\big),\;f\in\mathbb{H}(\beta,L),\;\theta\in\SSd\right\},
\end{equation}
$d\geq 2$ is the dimension and $\bob$ is the constant from Assumption \ref{ass2:assumption-on-kernel}, which can be arbitrary but must be chosen {\it a priory}.
\begin{theorem}
\label{th:pointwise-adaptation}
Let $\bob>0$ be fixed and let Assumptions \ref{ass:assumption-on-kernel} and \ref{ass2:assumption-on-kernel} hold. Then, for any
$\beta\leq\bob$, $L>0$ and $x\in[-1/2,1/2]^{2}$, we have
$$
\sup_{F\in\mathbb{F}_2(\beta,L)}\cc R_{r,x}^{(\eps)} \Big(\hat{F}_{(\hat{\theta}, \hat{h})}, F\Big)
\leq \|\K \|^2_{\infty}\left
C_{r,1}(Q,\K)\psi_\eps(\beta,L)
+C_{r,2}(L,Q,\K)\, \eps\sqrt{\ln (1/\eps)}\right],
$$
where $\psi_\eps(\beta,L)=L^{\frac{1}{2\beta+1} }\left(\eps\sqrt{\ln(1/\eps)}\right)^{\frac{2\beta}{2\beta+1}}$.
\smallskip
\par Moreover, for any $\beta,L>0$, \( r \ge 1 \), \( x \in [-1/2, 1/2]^d \) with \( d\ge 2 \) and any $\eps>0$ small enough,
$$
\inf_{\tilde{F}}\sup_{F\in\mathbb{F}_d(\beta,L)}\cc R_{r,x}^{(\eps)} \Big(\tilde{F}, F\Big)
\geq \kappa \psi_\eps(\beta,L),
$$
where infimum is over all possible estimators. Here $\kappa$ is a numerical constant independent of $\eps$ and $L$.
\end{theorem}
We conclude that the estimator $\hat{F}_{(\hat{\theta}, \hat{h})}$ is minimax adaptive with respect to the collection of classes
$\big\{\mathbb{F}_d(\beta,L),\;\; \beta\le \bob,\; L>0\big\}$. As already mentioned, this result is quite surprising. Indeed, if for example,
the directional vector $\theta=(1,0)^{\T}$, i.e. is known, then $\mathbb{F}(\beta,L)=\mathbb{H}(\beta,L)$ and the considered estimation problem
can be easily reduced to estimation of $f$ at a given point in the univariate Gaussian white noise model. As it is shown in \cite{Lep1990}
the adaptive estimator over the collection $\big\{\mathbb{H}(\beta,L),\;\; \beta\leq \bob,\; L>0\big\}$ does not exist.
Also, we would like to emphasize that the lower bound result given by the second assertion of the theorem is proved for arbitrary dimension. As to the proof of the first statement of the theorem it is based on the evaluation of the uniform, over $\mathbb{H}_d(\beta,L)$, lower bound for
$
h^*_{\K,f}(\cdot)
$
and on the application of Theorem \ref{th:local-oracle-inequality}. We note also that the upper bound for the minimax risk given in Theorem \ref{th:pointwise-adaptation} was earlier given in \cite{GoldLep2008}, but the estimation procedure used there is completely
different from our selection rule.
\subsection{Adaptive estimation under the \( L_r \) losses}\label{sect:Global_Adapt}
We start this section with the definition of the Nikol'skii class of functions.
\begin{definition}
\label{def:nikolskii-class}
Let \( \beta>0 \), \( L>0 \) and $p\in [1,\infty)$ be fixed. A function $g:\mathbb{R}\to\RR$ belongs to the Nikol'skii class $\mathbb{N}_p(\beta,L)$,
if $g$ is $m_\beta$-times continuously
differentiable and
\begin{eqnarray*}
&&\left( \int_{\RR}\left|g^{(m)}(t)\right|^p \dd t \right)^{\frac{1}{p} }\le L,\quad \forall m=1,\ldots, m_\beta;
\\
&&\left( \int_{\RR}\left|g^{(m_\beta)}(t+h)-g^{(m_\beta)}(t)\right|^p \dd z \right)^{\frac{1}{p} }
\leq Lh^{\beta-m_\beta},\;\;\forall h>0.
\end{eqnarray*}
\end{definition}
\noindent Later on we assume that $\mathbb{N}_p(\beta,L)=\mathbb{H}(\beta,L)$ if $p=\infty$.
\par Here the target of estimation is the entire function $F(\cdot)$ under the assumption that
$ F \in\mathbb{F}_p(\bob):=\bigcup_{\beta\leq\bob}\bigcup_{L>0}\mathbb{F}_{2,p}(\beta,L)$, where
\begin{equation}\label{Def:SI-class_global}
\mathbb{F}_{d,p}(\beta,L)=\left\{F:\RR^d\to\RR\;|\; F(z)=f\big(z^{\T}\theta\big),\;f\in\mathbb{N}_p(\beta,L),\;\theta\in\SSd\right\}.
\end{equation}
\begin{comment}
Let us briefly discuss the applicability of Theorem \ref{th:global-oracle-inequality} which requires
that $f\in\mathbb{F}_M$. In order to guarantee it we will assume that
$\beta p> 1$. The latter assumption is standard in estimation of functions possessing inhomogeneous smoothness, see for example,
\cite{DJKP}, \cite{LepMamSpok97}, \cite{KLP2008}.
If $\beta p> 1$ the embedding
$\mathbb{N}_p(\beta,L)\subset\mathbb{H}(\beta-1/p,cL)$ with some absolute constant $c>0$ guarantees that $f\in\mathbb{F}_{M},\; M=cL$
and Theorem \ref{th:global-oracle-inequality} is applicable.
\end{comment}
\begin{theorem}
\label{th:global-adaptation}
Let $\bob>0$ be fixed and let Assumptions \ref{ass:assumption-on-kernel} and \ref{ass2:assumption-on-kernel} hold. Then, for any
$L>0$, $p>1$, $p^{-1}<\beta\leq\bob$ and $r\geq 1$,
$$
\sup_{F\in\mathbb{F}_{2,p}(\beta,L)}\cc R_{r}^{(\eps)} \Big(\hat{F}_{(\hat{\theta}, \hat{h})}, F\Big)
\leq
\|\cc K\|^{2}_\infty \Big[ \kappa C_{r,1}(Q,\K)\varphi_\eps(\beta,L,p)+C_{r,2}(L,Q,\K)\eps\sqrt{\ln (1/\eps)}\Big],
$$
where
$$
\varphi_\eps(\beta,L,p)=\left\{
\begin{array}{lll}
L^{\frac{1}{2\beta+1}}\left(\eps\sqrt{\ln(1/\eps)}\right)^{\frac{2\beta}{2\beta+1}},\quad& (2\beta+1)p>r;
\\
L^{\frac{1}{2\beta+1}}\left(\eps\sqrt{\ln(1/\eps)}\right)^{\frac{2\beta}{2\beta+1}}\big[\ln(1/\eps)\big]^{\frac{1}{r}},\quad& (2\beta+1)p=r;
\\
L^{\frac{1/2-1/r}{\beta-1/p+1/2}}
\left(\eps\sqrt{\ln (1/\eps)}\right)^{\frac{\beta-1/p+1/r}{\beta-1/p+1/2}},\quad& (2\beta+1)p< r.
\end{array}
\right.
$$
The constant $\kappa$ is independent of $\eps$, $L$ and $\K$.
\end{theorem}
Let us make some remarks. First, note that $\mathbb{F}_{2,p}(\beta,L)\supset\mathbb{N}_{p}(\beta,L)$.
Indeed, the class $\mathbb{N}_{p}(\beta,L)$ can be viewed as the class of functions $F$ satisfying $F(\cdot)=f(\ta^\T\cdot)$ with
$\ta=(1,0)^{\T}$. Then, the problem of estimating such (2-variate) functions can be reduced to the estimation
of univariate functions observed
in the one-dimensional GWN model. In view of this remark the rate of convergence for the latter problem
(which can be found for example in \cite{DJKP,DelyonJuditski1996} ) is the lower bound for the minimax risk defined on
$\mathbb{F}_{2,p}(\beta,L)$. Under assumption $\beta p>1$ this rate of convergence is given by
$$
\phi_\eps(\beta,L,p)=\left\{
\begin{array}{lll}
L^{\frac{1}{2\beta+1}}\eps^{\frac{2\beta}{2\beta+1}},\quad& (2\beta+1)p> r;
\\
L^{\frac{1}{2\beta+1}}\left(\eps\sqrt{\ln(1/\eps)}\right)^{\frac{2\beta}{2\beta+1}},\quad& (2\beta+1)p=r;
\\
L^{\frac{1/2-1/r}{\beta-1/p+1/2}}
\left(\eps\sqrt{\ln (1/\eps)}\right)^{\frac{\beta-1/p+1/r}{\beta-1/p+1/2}},\quad& (2\beta+1)p< r.
\end{array}
\right.
$$
The minimax rate of convergence in the case $(2\beta+1)p=r$ remains an open problem, and the rate presented in the middle line above
is only the lower asymptotic bound for the minimax risk. Therefore the proposed estimator $\hat{F}_{(\hat{\theta}, \hat{h})}$
is adaptive whenever $(2\beta+1)p<r$.
In the case $(2\beta+1)p\geq r$ we loose only a logarithmic factor with respect to the optimal rate
and, as mentioned in Introduction, the construction of adaptive estimator over a collection
$\big\{\mathbb{F}_{2,p}(\beta,L), \;\beta>0,\;L>0\big\}$
in this case remains an open problem.
\section{Proofs}
\label{sec:proofs}
\subsection{Proof of Theorem \ref{th:local-oracle-inequality}}
The section starts with
the constants used in the statement of the theorem as well as technical lemmas whose proofs are postponed to Appendix.
\paragraph{Constants}
\begin{eqnarray*}
&&C_{r,1}(Q,\K)
=
8 \left[\Lambda(\K,Q)+\sqrt{4r+2}+1\right]
+
\mathfrak{c}_r \left[ (2 + \sqrt{2}) \Lambda(\K,Q) + 2 \right] +1;
\\*[2mm]
&&C_{r,2}(M,Q,\K)=2^{1/r}\left[2M+\Lambda(\K,Q) \mathfrak{c}_{2r} \right] .
\end{eqnarray*}
\subsubsection{Auxiliary results}
For any $\theta,\nu\in\SS$ and $h\in[\eps^2,1]$ denote
\begin{eqnarray*}
S_{(\ta, h)(\nu, h)}(x)
&=&
\det \big(E_{(\ta, h)(\nu,h)}\big)
\int K(E_{(\ta, h)(\nu,h)}(t-x))F(t)\dd t ,
\\
S_{(\ta, h)}(x)
&=&
\det \big(E_{(\ta, h)}\big)
\int K(E_{(\ta, h)}(t-x))F(t)\dd t .
\end{eqnarray*}
For ease of notation, we write $h^*_f=h^*_{\K,f}(x^{\T}\ta^{\circ})$.
\begin{lemma}
\label{lem:bounds for bias} Grant Assumption \ref{ass:assumption-on-kernel}. Then, for any $\nu\in\SS$ and any $\eta,h\in [\eps^2,1]$
satisfying $\eta\leq h\leq 2^{-1}h^*_{f}$, one has
\begin{eqnarray*}
&&\left|S_{(\ta^{\circ}, h)(\nu, h)}(x)-S_{(\nu, h)}(x)\right|\leq 2(h^*_f)^{-1/2}\| \K \|^{2}_{\infty}\;\eps \sqrt{\ln (1/\eps)};
\\[2mm]
&&\left|S_{(\nu, h)}(x)-S_{(\nu, \eta)}(x)\right|\leq 2(h^*_f)^{-1/2}\| \K \|^{2}_{\infty}\;\eps \sqrt{\ln (1/\eps)};
\\[2mm]
&&\left|S_{(\ta^{\circ}, h)} - F(x)\right|\leq (h^*_f)^{-1/2}\| \K \|_{\infty}\;\eps \sqrt{\ln (1/\eps)}.
\end{eqnarray*}
\end{lemma}
\smallskip
Let $\mathcal{E}_{a,A},\; 0<a,A<\infty,$ be a set of $2\times2$ matrices such that
$$
\left|\det(E)\right|\geq a,\quad |E|_\infty\leq A,\;\;\forall E\in\mathcal{E}_{a,A}.
$$
Here $|E|_\infty = \max_{i,j} |E_{i,j}|$ denotes the supremum norm, the maximum absolute value entry of the matrix \( E \). Later on without loss of generality we will assume that $a\leq A,\; A\geq 1$.
Assume that the function $ \cc L:\mathbb{R}^{2}\to\mathbb{R}$ is compactly supported on $[-1/2,1/2]^{2}$, $\int \cc L=1$ and satisfies the Lipschitz condition
$$
\left| \cc L(u)- \cc L(v)\right|\leq \Upsilon |u-v|_2,\;\;\forall u,v\in \RR^2,
$$
where $|\cdot|_2$ is the Euclidian norm. Let $y\in \RR^2$ be fixed. On the parameter set $\mathcal{E}_{a,A}$ let a Gaussian random function be defined by
$$
\zeta_{y}(E)= \| \cc L\|^{-1}_2\sqrt{\left|\det(E)\right|}\int \cc L\big(E(u-y)\big)W(\dd u).
$$
Put \( \mathbf{c}(a,A) =4\sqrt{2} \left[ \ln(A\vee\{A/a\}^2) + 2 \ln{(1+ \sqrt{2} \Upsilon)} \right]^{1/2} + 29 \)
and \( \mathfrak{c}_q=\left(\mathbb{E}\big(1+|\varsigma|\big)^{q}\right)^{1/q} \), where \( \varsigma\;\sim\;\norm{0}{1} \) .
\begin{lemma}
\label{lem:gauss-on-matrices}
For any \( z>0 \)
\begin{equation*}
\mathbb{P}\Big\{\sup_{E\in\mathcal{E}_{a,A}}\left|\zeta_{y}(E)\Big| \ge \mathbf{c}(a,A)+z\right\}\le \PP\{ |\varsigma| \ge z \}\le e^{-\frac{z^2}{2}}.
\end{equation*}
Moreover, for any \( q\ge 1 \)
\begin{equation*}
\left(\mathbb{E}\Big[\sup_{E\in\mathcal{E}_{a,A}}\big|\zeta_{y}(E)\big|\Big]^{q}\right)^{1/q}
\le
\mathfrak{c}_q \mathbf{c}(a,A)
\end{equation*}
\end{lemma}
\subsubsection{Proof of Theorem \ref{th:local-oracle-inequality}}
Let $h^*\in\mathcal{H_\eps}$ be such that
$
h^*\leq 2^{-1}h^*_{f}<2h^*
$.
Introduce the random events
$$
\cc A=\left\{(\ta^{\circ},h^*)\in\mathcal{P}(x)\right\},\quad
\cc B=\left\{\hat{h}\ge h^*\right\},\quad \cc C=\cc A \cap \cc B,
$$
and let \( \overline{\cc C} \) denote the event complimentary to $\mathcal{C}$. We split the proof into two steps.
\paragraph{Risk computation under \( \cc C \)} The triangle inequality gives
\begin{eqnarray}
\label{eq0:proof-of-theorem-local}
\left|\hat F_{(\hat{\ta},\hat{h})}(x)-F(x)\right|&\leq&\left|\hat F_{(\hat{\ta},\hat{h})}(x)- \hat F_{(\hat{\ta},h^*)}(x)\right|+
\left|\hat F_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(x)-\hat F_{(\hat{\ta},h^*)}(x) \right|
\nonumber\\*[2mm]
&\;&+\left|\hat F_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(x)-\hat F_{(\ta^{\circ},h^*)}(x) \right|+\left|\hat F_{(\ta^{\circ},h^*)}(x)-F(x)\right|.
\end{eqnarray}
$1^0.\;$ Since $h^*\geq 4^{-1}h^*_f$ the definition of $\hat{h}$ yields
\begin{eqnarray}
\label{eq1:proof-of-theorem-local}
\left|\hat F_{(\hat{\ta},\widehat{h})}(x)- \hat F_{(\hat{\ta},h^*)}(x)\right|\mathrm{1}_{\mathcal{B}}\leq \TH(h^*)\leq\TH\big(h_f^*/4\big).
\end{eqnarray}
Let us make some remarks. Note that $E_{(\theta, h)(\nu,h)}=\pm E_{(\nu, h)(\theta,h)}$ for any $\theta,\nu$ and $h$. Hence, we conclude that
$\hat F_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(\cdot)\equiv \hat F_{(\hat{\ta},h^*)(\ta^{\circ},h^*)}(\cdot)$ since $\mathcal{K}$ is symmetric, see Assumption \ref{ass:assumption-on-kernel}.
Next, we note that obviously $\mathcal{A}\subseteq\{\mathcal{P}(x)\neq\emptyset\}$
and, moreover, $\mathcal{A}\subseteq\big\{\tilde{h}\geq h^*\big\}$ in view of the definition
of $\tilde{h}$. Lastly, $\big(\hat{\theta},\tilde{h}\big)\in\cc P(x)$ by definition that means $R_{(\hat{\ta}, \tilde{h})}(x)\leq 0$. Consequently,
\begin{eqnarray}
\label{eq2:proof-of-theorem-local}
\hskip-1cm \left|\hat F_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(x)-
\hat F_{(\ta^{\circ},h^*)}(x) \right|\mathrm{1}_{\mathcal{A}}
&=&\left|\hat F_{(\hat{\ta},h^*)(\ta^{\circ},h^*)}(x)-\hat F_{(\ta^{\circ},h^*)}(x) \right|
\mathrm{1}_{\mathcal{A}}
\nonumber\\*[2mm]
&\leq&\TH(h^*) \leq \TH\big(h_f^*/4\big).
\end{eqnarray}
$2^0.\;$Introduce the following notations. For any $\ta,\nu\in\SS$ and $h\in[\eps^2,1]$ set
\begin{eqnarray*}
\xi _{(\ta, h)(\nu,h)}(x)
&=&
\|K\|^{-1}_2\sqrt{\det\big( E_{(\ta, h)(\nu,h)}\big) }
\int K\big( E_{(\ta,h)(\nu,h)}(t-x)\big)W (\dd t);
\\
\xi _{(\ta, h)}(x)
&=&
\|K\|^{-1}_2\sqrt{\det\big( E_{(\ta, h)}\big)}
\int K\big( E_{(\ta,h)}(t-x)\big)W (\dd t).
\end{eqnarray*}
We remark that $\big| E_{(\ta, h)}\big|_\infty\leq h^{-1}$ and $\big|E_{(\theta,h)(\nu,h)}\big|_\infty\leq h^{-1}$. Moreover,
\begin{eqnarray}
\label{eq3:proof-of-theorem-local}
&&(4h)^{-1}\leq\det\big(E_{(\theta,h)(\nu,h)}\big)\leq (2h)^{-1},\qquad\det\big( E_{(\ta, h)}\big)=h^{-1}.
\end{eqnarray}
Since $h\in[\eps^2,1]$, we assert that
\begin{equation}
\label{eq4:proof-of-theorem-local}
E_{(\theta,h)(\nu,h)}, E_{(\ta, h)}\in \mathcal{E}_{\frac{1}{4},\frac{1}{\eps^{2}}},\quad \forall \ta,\nu\in\SS,\;\forall h\in[\eps^2,1].
\end{equation}
We note also that
for any $\ta,\nu\in\SS$ and $h\in[\eps^2,1]$
\begin{eqnarray*}
&&\left|\hat F_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(x)-\hat F_{(\hat{\ta},h^*)}(x)\right|
\leq\left|S_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(x)-S_{(\hat{\ta},h^*)}(x)\right|
\\*[2mm]
&&+
\eps\|K\|_2\sqrt{\det\big( E_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}\big) }\left|\xi _{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(x)\right|
+\eps\|K\|_2\sqrt{\det\big( E_{(\hat{\ta},h^*)}\big) }\left|\xi _{(\hat{\ta},h^*)}(x)\right|.
\end{eqnarray*}
We obtain from the first assertion of Lemma \ref{lem:bounds for bias} with $\nu=\hat{\ta},\;h=h^*$,
(\ref{eq3:proof-of-theorem-local}) and (\ref{eq4:proof-of-theorem-local})
\begin{eqnarray}
\label{eq5:proof-of-theorem-local}
\left|\hat F_{(\ta^{\circ},h^*)(\hat{\ta},h^*)}(x)-\hat F_{(\hat{\ta},h^*)}(x)\right|
&\leq&
\frac{2\,\| \K \|^{2}_{\infty}}{\sqrt{h^*_f}} \,\eps \sqrt{\ln (1/\eps)}
+
\frac{2 + \sqrt{2}}{\sqrt{h^*_f}}\;\| \K \|^{2}_{2}\,\eps \sqrt{\ln (1/\eps)}\;\zeta_\eps(x)
\nonumber\\
&\leq&
\frac{\| \K \|^{2}_{\infty}}{\sqrt{h^*_f}} \, \eps \sqrt{\ln (1/\eps)}\,\big[2+(2+\sqrt{2})\zeta_\eps(x)\big],
\end{eqnarray}
where we denoted
$$
\zeta_\eps=\left[\ln{(1/\eps)}\right]^{-1/2}
\displaystyle{\sup_{E\in \mathcal{E}_{\frac{1}{4},\frac{1}{\eps^{2}}}}}\left|\zeta_x\big(E\big)\right|.
$$
We have also used that $2h^*\leq h^*_f<4h^*$.
\smallskip
$3^0.\;$ We get in view of the third assertion of Lemma \ref{lem:bounds for bias} that
\begin{eqnarray}
\label{eq6:proof-of-theorem-local}
\left|\hat F_{(\ta^{\circ},h^*)}(x)-F(x)\right|
&\leq& \sqrt{1/h_f^*}\;\| \K \|_{\infty}\;\eps
\sqrt{\ln (1/\eps)}+\sqrt{1/h^*}\;\| \K \|^{2}_{2}\;\eps|\varsigma|
\nonumber\\*[2mm]
&\leq&\sqrt{1/h_f^*}\;\| \K \|^2_{\infty}\;\eps \sqrt{\ln (1/\eps)}\big(1+2|\varsigma|\big),
\end{eqnarray}
where $\varsigma\;\sim\;\mathcal{N}(0,1).$
\smallskip
$4^0.\;$ We obtain from (\ref{eq0:proof-of-theorem-local}), (\ref{eq1:proof-of-theorem-local}), (\ref{eq2:proof-of-theorem-local}), (\ref{eq5:proof-of-theorem-local}) and (\ref{eq6:proof-of-theorem-local}) and the second assertion of Lemma \ref{lem:gauss-on-matrices}
with $ \cc L=K,\;a=1/4,\; A=\eps^{-2}$ and $q=r$, noting that $\Upsilon=\sqrt{2}Q\|\K\|_\infty$ ,
\begin{eqnarray}
\label{eq7:proof-of-theorem-local}
\left\{\mathbb{E}\left|\hat F_{(\hat{\ta},\hat{h})}(x)-F(x)\right|^r\mathrm{1}_{\mathcal{C}}\right\}^{1/r}
&\le&
2\TH\big(h^*_f/4\big)
+
\left[ (2 + \sqrt{2}) \Lambda(\K,Q) \mathfrak{c}_r + 2 \mathfrak{c}_r +1 \right]
\frac{\| \K \|^{2}_{\infty}}{\sqrt{h^*_f}}\eps \sqrt{\ln (1/\eps)}
\nonumber\\
&\le&
C_{r,1} \frac{\| \K \|^{2}_{\infty}}{\sqrt{h^*_f}} \eps \sqrt{\ln (1/\eps)},
\end{eqnarray}
Here we have also used that
$$
\sup_{\eps\leq e^{-1}}
\left[
\frac{4\sqrt{2} \left\{ 2 \sqrt{\ln{(2/\eps)}}+ \sqrt{2\ln(1+2Q\|\K\|_\infty)} \right\} + 29}{\sqrt{\ln{(1/\eps)}}}
\right]\leq \Lambda(\K,Q).
$$
\paragraph{Risk computation under \( \overline{\cc C} \)}
Since $f\in\mathbb{F}_M$ one can easily evaluate the discrepancy between the adaptive estimator and the value of function
$$
\left|\hat F_{(\hat{\ta},\hat{h})}(x)-F(x)\right|\leq M\big(1+\|K\|_1\big)+\eps\|K\|_2\sqrt{\det\big( E_{(\hat{\ta},\hat{h})}\big) }\left|\xi _{(\hat{\ta},\hat{h})}(x)\right|.
$$
We obtain in view of (\ref{eq3:proof-of-theorem-local}) and (\ref{eq4:proof-of-theorem-local}), taking into account that $\hat{h}>\eps^{2}$,
$$
\left|\hat F_{(\hat{\ta},\hat{h})}(x)-F(x)\right|\leq M\big(1+\|\K\|^{2}_1\big)+\|\K\|^2_2\sqrt{\ln (1/\eps)}\zeta_\eps.
$$
Thus, applying the second assertion of Lemma \ref{lem:gauss-on-matrices}
with $ \cc L=K,\;a=1/4,\; A=\eps^{-2}$, $\Upsilon=\sqrt{2}Q\|\K\|_\infty$ and $q=2r$, we get
\begin{equation*}
\left[\mathbb{E}_F^{\eps}\left|\hat F_{(\hat{\ta},\hat{h})}(x)-F(x)\right|^{2r}\right]^{1/2r}
\le
\left[2M+\Lambda(\K,Q) \mathfrak{c}_{2r} \right] \|\K\|^2_\infty \sqrt{\ln (1/\eps)}.
\end{equation*}
Here it is used that $1\le\|\K\|_1\le \|\K\|_2\le \|\K\|_\infty$ due to
Assumption \ref{ass:assumption-on-kernel} ({\it 1}) and that $\eps\leq e^{-1}$.
\smallskip
With $\lambda_r(M,\K,Q)=2M+\Lambda(\K,Q) \mathfrak{c}_{2r}$ the use of the Cauchy-Schwartz inequality leads to the following bound:
\begin{eqnarray}
\label{eq8:proof-of-theorem-local}
&&\left\{\mathbb{E}_F^\eps\left|\hat F_{(\hat{\ta},\hat{h})}(x)-F(x)\right|^r\mathrm{1}_{\overline{\mathcal{C}}}\right\}^{1/r}
\lambda_r(M,\K,Q) \| \K \|^2_{\infty}\sqrt{\ln (1/\eps)}
\left[\mathbb{P}_F^\eps(\overline{\mathcal{A}})+\mathbb{P}_F^\eps(\overline{\mathcal{B}})\right]^{1/2r}.
\end{eqnarray}
$1^0.\;$ Let us bound from above $\mathbb{P}_F^\eps(\overline{\mathcal{A}})$. We note that
\begin{eqnarray}
\label{eq9:proof-of-theorem-local}
\mathbb{P}_F^\eps(\overline{\mathcal{A}})
&=&\mathbb{P}_F^\eps\big\{(\ta^{\circ},h^*)\notin \cc P(x)\big\}
= \mathbb{P}_F^\eps\big\{R_{(\ta^{\circ}, h^*)}(x)>0\big\}\nn*[2mm]
&\leq&\sum_{k:\;\eps^2\leq 2^{-k}\leq h^*}\mathbb{P}_F^\eps\left\{\sup_{\nu\in\SS}
\big|\hat F_{(\ta^{\circ}, 2^{-k})(\nu, 2^{-k})}(x) - \hat F_{(\nu, 2^{-k})}(x) \big|> \TH\big(2^{-k}\big)\right\}.
\end{eqnarray}
For any $k$ satisfying $2^{-k}\leq h^*$ and any $\nu\in\SS$, similarly to (\ref{eq5:proof-of-theorem-local}), we obtain from the first assertion of Lemma \ref{lem:bounds for bias}
with $h=2^{-k}$, (\ref{eq3:proof-of-theorem-local}) and (\ref{eq4:proof-of-theorem-local}) that
\begin{eqnarray}
\label{eq10:proof-of-theorem-local}
\hskip-0.7cm\left|\hat F_{(\ta^{\circ},2^{-k})(\nu,2^{-k})}(x)-\hat F_{(\nu,2^{-k})}(x)\right|
&\le& 2(h^*_f)^{-1/2}
\| \K \|^{2}_{\infty}\;\eps \sqrt{\ln (1/\eps)}+2\sqrt{2^k}\;\| \K \|^{2}_{2}\eps \sqrt{\ln (1/\eps)}\;\zeta_\eps(x)
\nonumber\\*[2mm]
&\le& 2^{1+k/2}\;\| \K \|^{2}_{\infty}\eps \sqrt{\ln (1/\eps)}\big[1+\zeta_\eps(x)\big].
\end{eqnarray}
Here we have also used that $h^*_f\geq 2^{-k}$. Remembering, that
$$
\TH(\eta)=2\| \K \|^{2}_{\infty}\left[\Lambda(\K,Q)+\sqrt{4r+2}+1\right]\eps\sqrt{\eta^{-1}\ln (1/\eps)},
$$
we obtain from (\ref{eq10:proof-of-theorem-local}) for any $k$ satisfying $2^{-k}\leq h^*$
\begin{eqnarray*}
&&\mathbb{P}_F^\eps\left\{\sup_{\nu\in\SS}
\big|\hat F_{(\ta^{\circ}, 2^{-k})(\nu, 2^{-k})}(x) - \hat F_{(\nu, 2^{-k})}(x) \big|> \TH\big(2^{-k}\big)\right\}
\nonumber
\\*[2mm]
&&\leq \mathbb{P}_F^\eps\bigg\{\displaystyle{\sup_{E\in \mathcal{E}_{\frac{1}{4},\frac{1}{\eps^{2}}}}}\left|\zeta_x\big(E\big)\right|
> \mathbf{c}\big(1/4,\eps^{-2}\big)+ \sqrt{(4r+2)\ln (1/\eps)}\bigg\}\le \eps^{2r+1},
\end{eqnarray*}
in view of the first assertion of Lemma \ref{lem:gauss-on-matrices}. It yields, together with (\ref{eq9:proof-of-theorem-local})
\begin{eqnarray}
\label{eq11:proof-of-theorem-local}
\mathbb{P}_F^\eps(\overline{\mathcal{A}})\le 2 \eps^{2r+1}\log_2 (1/\eps) \le 2 \eps^{2r}.
\end{eqnarray}
$2^0.\;$ An upper bound on the probability of event \( \left\{\hat{h}< h^*\right\} \) is given by
\begin{eqnarray}
\label{eq12:proof-of-theorem-local}
\mathbb{P}_F^\eps(\overline{\mathcal{B}})&=&\mathbb{P}_F^\eps\left(\bigcup_{k:\;\eps^2\leq 2^{-k}\leq h^*}
\left\{\left|\hat F_{(\hat{\ta},h^*)}(x)- \hat F_{(\hat{\ta}, 2^{-k})}(x)\right|
>\TH\big(2^{-k}\big)\right\}\right)=
\nonumber\\*[2mm]
&\leq&\sum_{k:\;\eps^2\leq 2^{-k}\leq h^*}\mathbb{P}_F^\eps
\left\{\left|\hat F_{(\hat{\ta},h^*)}(x)- \hat F_{(\hat{\ta}, 2^{-k})}(x)\right| > \TH\big(2^{-k}\big)\right\}.
\end{eqnarray}
We note that
\begin{eqnarray*}
&&\left|\hat F_{(\hat{\ta},h^*)}(x)-\hat F_{(\hat{\ta},2^{-k})}(x)\right|
\leq \left|S_{(\hat{\ta},h^*)}(x)-S_{(\hat{\ta},2^{-k})}(x)\right|
\\*[2mm]
&&+
\eps\|K\|_2\sqrt{\det\big( E_{(\hat{\ta},h^*)}\big) }\left|\xi _{(\hat{\ta},h^*)}(x)\right|
+\eps\|K\|_2\sqrt{\det\big( E_{(\hat{\ta},2^{-k})}\big) }\left|\xi _{(\hat{\ta},2^{-k})}(x)\right|.
\end{eqnarray*}
Applying the second assertion of Lemma \ref{lem:bounds for bias}
with $\nu=\hat{\ta},\;h=h^*,\;\eta=2^{-k}$, (\ref{eq3:proof-of-theorem-local}) and (\ref{eq4:proof-of-theorem-local})
\begin{eqnarray}
\label{eq13:proof-of-theorem-local}
\left|\hat F_{(\hat{\ta},h^*)}(x)-\hat F_{(\hat{\ta},2^{-k})}(x)\right|
&\leq& 2(h^*_f)^{-1/2}\; \| \K \|^{2}_{\infty}\;\eps \sqrt{\ln (1/\eps)}+2\sqrt{2^k}\;\| \K \|^{2}_{2}\eps \sqrt{\ln (1/\eps)}\;\zeta_\eps(x)
\nn *[2mm]
&\leq& 2^{1+k/2}\| \K \|^{2}_{\infty}\eps \sqrt{\ln (1/\eps)}\big[1+\zeta_\eps(x)\big].
\end{eqnarray}
We remark that the right-hand sides of (\ref{eq10:proof-of-theorem-local}) and (\ref{eq13:proof-of-theorem-local}) coincide and, therefore,
repeating the computation led to (\ref{eq11:proof-of-theorem-local}) we get
\begin{eqnarray}
\label{eq14:proof-of-theorem-local}
\mathbb{P}_F^\eps(\overline{\mathcal{B}})\le 2 \eps^{2r}.
\end{eqnarray}
We obtain from (\ref{eq8:proof-of-theorem-local}), (\ref{eq11:proof-of-theorem-local}) and (\ref{eq14:proof-of-theorem-local})
\begin{eqnarray}
\label{eq15:proof-of-theorem-local}
\left\{\mathbb{E}_F^\eps\left|\hat F_{(\hat{\ta},\hat{h})}(x)-F(x)\right|^r\mathrm{1}_{\overline{\mathcal{C}}}\right\}^{1/r}
&\leq& 2^{1/r}\lambda_r(M,\K,Q) \| \K \|^2_{\infty}\eps\sqrt{\ln (1/\eps)}.
\end{eqnarray}
The assertion of the theorem follows now from (\ref{eq7:proof-of-theorem-local}) and (\ref{eq15:proof-of-theorem-local}).
\epr
\subsection{Proof of Theorem \ref{th:pointwise-adaptation}}
We start this section with an auxiliary result used in the proof of the second assertion of the theorem. That result is proved in \cite{KLP2008}, Proposition 7, and for convenience, we formulate it as Lemma \ref{lem:KLP-result} below.
\subsubsection{Auxiliary result}
The result cited below concerns a lower bound for estimators of an arbitrary mapping in the framework of GWN model. Below a version adjusted to the estimation at a given point is provided.
\smallskip
Let \( \cc F \) be a nonempty class of functions and let \( F: \RR^d \to \RR \) be an unknown signal from model \eqref{WGN_model}--\eqref{single-index} satisfying \( F \in \mathcal{F} \subset \mathbb{L}_2(\cc D) \), \( \cc D =[-1, 1]^d \). The aim is to estimate the functional \( F(x) \), \( x \in [-1/2, 1/2]^d \).
\begin{lemma}
\label{lem:KLP-result} (\cite{KLP2008})
Assume that for any \( \eps > 0 \) there exist a positive integer \( \numhyp \), \( \numhyp \to \infty \) as \( \eps \to 0 \), \( \rho \in (0,1) \),
\( c>0 \) and functions \( F_0, F_1, \ldots , F_{\numhyp} \in \cc F \) such that:
\begin{eqnarray}
\label{eq:ass1-klp-lemma}
&&|F_i(x) - F_0(x)| =\lambda_\eps,\qquad\; \forall i=1, \ldots, \numhyp;
\\*[2mm]
\label{eq:ass2-klp-lemma}
&& \langle F_i - F_0 , F_j - F_0 \rangle \le c \eps^2\quad\;\;
\forall i,j=1, \ldots, \numhyp,\; i \ne j;
\\*[2mm]
\label{eq:ass3-klp-lemma}
&&\| F_i - F_0 \|^2_2 \le \rho \eps^2 \ln (\numhyp),\quad
\forall i=1, \ldots, \numhyp.
\end{eqnarray}
Then for \( r\geq 1 \)
\begin{equation*}
\label{lower bound GWN}
\inf_{\tilde F}
\sup_{F \in \cc F}
\left(\expect{F}
\big| \tilde{F} (x) - F (x) \big|^r\right)^{\frac{1}{r}}
\ge
\frac{1}{2}\left( 1- \sqrt{\frac{e^c -1}{e^c +3} } \right) \lambda_\eps.
\end{equation*}
\end{lemma}
\subsubsection{Proof of Theorem \ref{th:pointwise-adaptation}}
\paragraph{Proof of the first assertion}
Under Assumptions \ref{ass:assumption-on-kernel} and \ref{ass2:assumption-on-kernel} the standard computation of the bias of kernel estimators, for any $f\in\mathbb{H}(\beta,L)$ and any \( z\in\RR \), gives
\begin{equation*}
\Delta_{\K,f}(h,z)
\leq
\frac{ L h^\beta2^{-\beta} \|\K\|_\infty }{ (1+\beta)m_{\beta}! }
\leq
\|\K\|_\infty Lh^\beta.
\end{equation*}
The right-hand side of the latter inequality does not depend of \( z \) so
\begin{equation*}
\Delta^*_{\K,f}(h,z)
\leq
\|\K\|_\infty Lh^\beta.
\end{equation*}
Hence,
$h^*_{\K,f}(z)\geq \left(L^{-1}\eps\sqrt{\ln (1/\eps)}\right)^{2/(2\beta+1)}$ for any \( z\in\RR \) and the first assertion of the theorem
follows from Theorem \ref{th:local-oracle-inequality}.
\paragraph{Proof of the second assertion}
The proof is based on the construction of a family $F_0, \ldots,F_{\numhyp}\in \cc F = \mathbb{F}_d(\beta,L) \subset L_2([-1,1]^d)$ satisfying conditions
\eqref{eq:ass1-klp-lemma}--\eqref{eq:ass3-klp-lemma} of Lemma \ref{lem:KLP-result}.
\smallskip
$1^0.\; $ Firstly, we construct \( F_0,\ldots,F_{\numhyp} \) and verify \eqref{eq:ass1-klp-lemma}.
Let \( g:\RR \to \RR \) be such that
\( \supp (g) \subset (-1/2, 1/2) \), \( g \in \mathbb{H}(\beta,1) \
and $g(0)\neq 0$.
Put \( h=\left(\mathfrak{a}L^{-1}\eps\sqrt{\ln (1/\eps)}\right)^{2/(2\beta+1)} \), where the constant $\mathfrak{a}>0$
will be chosen later in order to satisfy \eqref{eq:ass3-klp-lemma}.
For any fixed \( u \in \RR \) define
\begin{equation}\label{link function for hypothesis}
f_u(v) = Lh^\beta g\big[(v-u) h^{-1}\big] \,, \;\; v\in\RR.
\end{equation}
For \( b>0 \) put \( \numhyp = \eps^{-b} \) assuming without loss of generality that \( \numhyp \) is an integer.
The value of \( b \) will be determined later in order to satisfy \eqref{eq:ass2-klp-lemma}.
Let $\left\{\vartheta_i,\; i=1,\ldots, \numhyp \right\}\subset \SSd$ be defined as follows:
\begin{equation*}
\vartheta_i=\big(\theta^{(1)}_i,\theta^{(2)}_i,0,\ldots,0\big)^{\T},
\qquad
\theta^{(1)}_i=\cos(i/\numhyp),
\quad
\theta^{(2)}_i=\sin(i/\numhyp).
\end{equation*}
Finally, set
\begin{equation}\label{hypotheses}
F_0 \equiv 0 \;\; \text{ and }\;\; F_i(t) = f_{\vartheta_i^{\T} x}\big(\vartheta_i^{\T} t\big) \;,\;\; i = 1, \ldots, \numhyp.
\end{equation}
As \( g \in \mathbb{H}(\beta,1) \) so \( f_u \) defined by \eqref{link function for hypothesis} belongs to \( \mathbb{H}(\beta,L) \) for any \( u \in \RR \) and therefore
all \( F_i \) are in \( \cc F= \mathbb{F}_d(\beta,L) \). Moreover, for any $ i=1,\ldots, N_\eps$
\begin{eqnarray}
\label{eq:check-first-condition}
\big|F_i(x) - F_0(x)\big|
=
\big|f_{\vartheta_i^{\T} x}\big(\vartheta_i^{\T} x\big)\big|
&=&
|g(0)| L^{\frac{1}{2\beta+1}}
\left(\mathfrak{a}\eps\sqrt{\ln (1/\eps)}\right)^{\frac{2\beta}{2\beta+1}}\nn
&=&
|g(0)| \mathfrak{a}^{\frac{2\beta}{2\beta+1}}\psi_\eps(\beta,L).
\end{eqnarray}
We see that \eqref{eq:ass1-klp-lemma} holds with
\( \lambda_\eps = |g(0)| \mathfrak{a}^{\frac{2\beta}{2\beta+1}}\psi_\eps(\beta,L) \).
\smallskip
$2^0.\; $ Now we check (\ref{eq:ass2-klp-lemma}).
Set \( \ort{\theta_{i}}=(-\sin(i/\numhyp)),\cos(i/\numhyp) \). We have
\begin{eqnarray*}
\langle F_i, F_j \rangle
&=&
L^2 h^{2\beta}
\int_{[-1 , 1 ]^d}
g\big(h^{-1}\vartheta_i^{\T}(t-x)\big)
g\big(h^{-1}\vartheta_j^{\T}(t-x)\big)\dd t \\
&\le&
3^{d-2}L^2 h^{2\beta+2} \int_{\mathbb{R}^2}\left|g\big(\theta_i^{\T}u\big)
g\big(\theta_j^{\T}u\big)\right|\dd u
=
3^{d-2}L^2 h^{2\beta+2} \big|\ort{\ta_{i}}^{\T} \ta_j\big|^{-1}\| g \|_{1}^2.
\\
&=&
3^{d-2}L^2 h^{2\beta+2} \big|\cos(j/\numhyp)\sin(i/\numhyp)-\cos(i/\numhyp)\sin(j/\numhyp)\big|^{-1}\| g \|_{1}^2
\\*[2mm]
&=&
3^{d-2}L^2 h^{2\beta+2} \big|\sin\big((i-j)/\numhyp\big)\big|^{-1}\| g \|_{1}^2
\\*[2mm]
&=&3^{d-2}L^2 h^{2\beta+2}\big(\sin\big(|i-j|/\numhyp\big)\big)^{-1}\| g \|_{1}^2.
\end{eqnarray*}
Thus, we obtain
\begin{eqnarray}
\label{eq1:proof-the-local-adaptive}
\sup_{i \ne j ; \;i,j = 1, \ldots, \numhyp} \langle F_i, F_j \rangle
& \le &
3^{d-2} L^2 h^{2\beta+2} \big(\sin\big(1/\numhyp\big)\big)^{-1}\| g \|_{1}^2
\le 3^{d-2} 2L^2 h^{2\beta+2}\numhyp \| g \|_{1}^2
\nonumber\\
&=& 3^{d-2} 2 \| g \|_{1}^2\mathfrak{a}^2\eps^{2}\ln (1/\eps)[\numhyp h].
\end{eqnarray}
Hence, choosing $b<2/(2\beta+1)$ we conclude that (\ref{eq:ass2-klp-lemma}) holds with any given $c>0$
for \( \eps \) small enough.
\smallskip
$3^0.\; $ It remains to verify \eqref{eq:ass3-klp-lemma}. By changing variables for any $i=1,\ldots, N_\eps$
\begin{eqnarray*}
\| F_i \|_2^2
& \le &
3^{d-1} \| g\|_2^2 L^2 h^{2\beta+1}
=
3^{d-1} \| g \|_{2}^2\mathfrak{a}^2 \eps^{2}\ln (1/\eps)
=
3^{d-1} \| g \|_{2}^2 \mathfrak{a}^2 b^{-1} \eps^{2}\ln \big(N_\eps\big).
\end{eqnarray*}
Here the notation \( \| \cdot \|_2 \) stands for the \( L_2 \) norms on \( [-1, 1]^d \) and \( [-1/2, 1/2] \) correspondingly. Choosing $\mathfrak{a}^2= 3^{- d}b\, \| g \|_{2}^{-2}$ we see that (\ref{eq:ass3-klp-lemma}) is fulfilled with $\rho=1/3$.
In view of (\ref{eq1:proof-the-local-adaptive}) the constants $c$ from (\ref{eq:ass2-klp-lemma}) and $\mathfrak{a}$ are chosen independently of $L$. Thus, the second assertion of the theorem follows from Lemma \ref{lem:KLP-result}.
\epr
\subsubsection{Proof of Theorem \ref{th:global-adaptation}}
To prove the theorem we will exploit the ideas developed in \cite{LepMamSpok97}. Moreover, our considerations are, to a great degree, based on the technical result of Lemma \ref{lem:l_p-norm-of-bais} below. Its proof is postponed until Appendix.
\begin{lemma}
\label{lem:l_p-norm-of-bais}
Grant Assumptions \ref{ass:assumption-on-kernel} and \ref{ass2:assumption-on-kernel}. Then, for any $\mathfrak{p}> 1$,
$0<s\leq \bob$, $\cQ>0$,
$$
\sup_{g\in\mathbb{N}_\mathfrak{p}(s,\cQ)}
\left\|\Delta^*_{\K,g}(h,\cdot)\right\|_\mathfrak{p}\leq 2\cQ h^{s}\|\K\|_\infty
(\tau_\mathfrak{p}+1)\left[2^{s \mathfrak{p}}-1\right]^{-\frac{1}{\mathfrak{p}}},\;\;\forall h>0.
$$
Here $\tau_\mathfrak{p}$ is a depending only of \( \mathfrak{p} \) constant from the $(\mathfrak{p},\mathfrak{p})$-strong maximal inequality.
\end{lemma}
\paragraph{Proof of Theorem \ref{th:global-adaptation}}
It is suffice to prove the theorem only in the case $r\geq p$. Indeed, remind that the risk $\cc R_{r}^{(\eps)}(\cdot, \cdot)$
is described by the \( L_r \) norm on $[-1/2,1/2]$, therefore
$$
\cc R_{r}^{(\eps)}(\cdot, \cdot)\leq \cc R_{p}^{(\eps)}(\cdot, \cdot), \;\;
r\leq p.
$$
Hence the case $r\leq p$ can be reduced to the case $r=p$.
Yet another observation. In view of embedding of Nikol'skii class \( \mathbb{N}_p(\beta,L) \) in the H\"{o}lder class with parameters \( \beta -1/p \) and \( c L \), \( c>0 \), the assumption \( \beta p > 1 \) provides that
\( f\in\mathbb{F}_M \) and the assumptions of Theorem \ref{th:global-oracle-inequality} are fulfilled. Moreover, in order to obtain the desired the assertion it suffices to bound from above
$
\left\| \sqrt{\frac{\| \K \|^2_{\infty}\eps^{2} \ln (1/\eps)}{h^*_{\K,f}(\cdot)}}\right\|_r.
$
\smallskip
Set $\Gamma_0=\big\{y\in[-1/2,1/2]: \;\; h^*_{\K,f}(y)=1\big\}$ and
$\Gamma_k=\big\{y \in [-1/2,1/2]: \;\; h^*_{\K,f}(y) \in (2^{-k}, 2^{-k+1}] \cap[\eps^2,1]\big\}$ for \( k=1,2,\ldots \; \). Later on, the integration over empty set is supposed to be zero.
We have
\begin{eqnarray*}
&& \left\|\sqrt{\frac{\|\K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(\cdot)}}\right\|^r_r
= \sum_{k\geq 1}\int_{\Gamma_k}\left(\sqrt{\frac{\|\K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(y)}}\right)^r\dd y+\int_{\Gamma_0}\left(\sqrt{\frac{\|\K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(y)}}\right)^r\dd y.
\end{eqnarray*}
The definition of $\Gamma_0$ implies
\begin{equation}
\label{eq2:proof-theorem-global-adaptation}
\int_{\Gamma_0}\left|\sqrt{\frac{\|\K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(y)}}\right|^r\dd t\leq \left[\|\K \|^2_{\infty}\eps^{2}\ln (1/\eps)\right]^{\frac{r}{2}}.
\end{equation}
Assumption
\ref{ass:assumption-on-kernel} ({\it 2}) implies that $\Delta^{*}_{\K,f}(\cdot,y)$ is continuous on $[\eps^2, 1]$, hence for any $k\geq 1$
\begin{equation}
\label{eq3:proof-theorem-global-adaptation}
\Delta^*_{\K,f}\big( h^*_{\K,f}(y),y\big)=\left[\frac{\| \K \|^2_{\infty}\eps^2 \ln (1/\eps)}{ h^*_{\K,f}(y)}\right]^{\frac{1}{2}},
\;\;\forall y\in\Gamma_k.
\end{equation}
Let $0\leq q_k\leq r$ be a sequence whose choice will be done later. We obtain from (\ref{eq3:proof-theorem-global-adaptation})
\begin{eqnarray}
\label{eq4:proof-theorem-global-adaptation}
\sum_{k\geq 1}\int_{\Gamma_k}\left(\sqrt{\frac{\| \K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{h^*_{\K,f}(y)}}\right)^r\dd y
&\le& \sum_{k\geq 1}\left(\frac{\| \K \|^2_{\infty}\eps^{2} \ln (1/\eps)}{2^{-k}}\right)^{\frac{r-q_k}{2}}
\int_{\Gamma_k}\left(\Delta^*_{\K,f}\big(2^{1-k} ,y\big)\right)^{q_k}\dd y \nn
&\le& \sum_{k\geq 1}\left(\frac{\| \K \|^2_{\infty}\eps^{2} \ln (1/\eps)}{2^{-k}}\right)^{\frac{r-q_k}{2}}
\int\left( \Delta^*_{\K,f}\big(2^{1-k} ,y\big)\right)^{q_k}\dd y=:\Xi.
\end{eqnarray}
To get the first inequality we have used that $\Delta^*_{\K,f}\big(\cdot,y\big)$ in monotonically increasing function.
\par The computation of the quantity on the right-hand side of (\ref{eq4:proof-theorem-global-adaptation}), including the choice of
$(q_k,\;k\geq 1)$,
will be done differently in dependence on
$\beta, p$ and $r$. Later on $c_1,c_2,\ldots,$ denote constants independent on $\eps$, $L$ and $\K$.
\smallskip
$1^0.\;$ {\it Case $(2\beta+1)p > r$.} Put
$$
h^*=\left[L^{-2}\eps^{2}
\ln (1/\eps)\right]^{\frac{1}{2\beta+1}}
$$
and choose $q_k=p$ if $2^{-k}\leq h^*$ and $q_k=0$ if $2^{-k}> h^*$.
Applying Lemma \ref{lem:l_p-norm-of-bais} with $\mathfrak{p}=p$, $s=\beta$ and $\cQ=L$ we get
\begin{eqnarray}
\label{eq55:proof-theorem-global-adaptation}
\Xi
&\le& c_1\big(L\|\K\|_\infty\big)^{p}\sum_{k:\;2^{-k}\leq h^*}\left(\frac{\| \K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{2^{-k}}\right)^{\frac{r-p}{2}}2^{-k\beta p}+c_2\left(\frac{\| \K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{h^*}\right)^{\frac{r}{2}} \nonumber\\
&\le& c_3 \|\K\|^{r}_\infty\left[L^{p}\left(\eps^{2}
\ln (1/\eps)\right)^{\frac{r- p}{2}}\sum_{k:\;2^{-k}\leq h^*}2^{-k\big[\beta p-\frac{r-p}{2}\big]}
+\left(\frac{\eps^{2}
\ln (1/\eps)}{h^*}\right)^{\frac{r}{2}}\right].
\end{eqnarray}
Because in the considered case $\beta p-\frac{r-p}{2}>0$, we obtain
\begin{eqnarray*}
\Xi\leq c_4 \|\K\|^{r}_\infty\left[L^{ p}\left(\eps^{2}
\ln (1/\eps)\right)^{\frac{r- p}{2}}(h^*)^{\beta p-\frac{r- p}{2}}
+\left(\frac{\eps^{2}
\ln (1/\eps)}{h^*}\right)^{\frac{r}{2}}\right].
\end{eqnarray*}
It remains to note that $h^*$ is chosen by balancing two terms on the right-hand side of the latter inequality. It yields
\begin{eqnarray}
\label{eq5:proof-theorem-global-adaptation}
\Xi\leq 2c_4 \left[\| \K \|_{\infty}L^{\frac{1}{2\beta+1}}\left(\eps\sqrt{\ln(1/\eps)}\right)^{\frac{2\beta}{2\beta+1}}\right]^r.
\end{eqnarray}
The argument in the case $(2\beta+1)p> r$ is completed with the use of Theorem~\ref{th:global-oracle-inequality},
(\ref{eq2:proof-theorem-global-adaptation}) and (\ref{eq5:proof-theorem-global-adaptation}).
\medskip
$2^0.\;$ {\it Case $(2\beta+1)p=r$.} Put $h^*=1$ and choose $q_k=p$ for all $k\geq 1$. Repeating the computations
led to (\ref{eq55:proof-theorem-global-adaptation}) we get
\begin{eqnarray}
\label{eq6:proof-theorem-global-adaptation}
\Xi\leq c_5\ln(1/\eps)\left[\|\K\|_\infty L^{ p/r}\left(\eps^{2}
\ln (1/\eps)\right)^{\frac{r-p}{2r}}\right]^r.
\end{eqnarray}
Here we have used that $\beta p-\frac{r-p}{2}=0$ and that the summation in (\ref{eq4:proof-theorem-global-adaptation})
runs over $k$ such that $2^{-k}\geq \eps^{2}$, since otherwise $\Gamma_k=\emptyset$. It remains to note that the equality $(2\beta+1)p=r$
is equivalent to $p/r=1/(2\beta+1)$ and $(r-p)/2r=\beta/(2\beta+1)$. The assertion of the theorem in the case $(2\beta+1)p=r$
follows now from Theorem \ref{th:global-oracle-inequality},
(\ref{eq2:proof-theorem-global-adaptation}) and (\ref{eq6:proof-theorem-global-adaptation}).
\medskip
$3^0.\;$ {\it Case $(2\beta+1)p<r$.} Choose $q_k=r$ if $2^{-k}\leq h^*$ and $q_k=p$ if $2^{-k}> h^*$,
where the choice of $h^*$ will be done later.
The following embedding holds, see \cite{BesovIlNik}:
$\mathbb{N}_p(\beta,L)\subseteq\mathbb{N}_r\big(\beta-1/p+1/r,c_6L\big)$. Thus, applying Lemma \ref{lem:l_p-norm-of-bais}
with $\mathfrak{p}=r$, $s=\beta-1/p+1/r$ and $\cQ=c_6L$ we get
\begin{eqnarray}
\label{eq7:proof-theorem-global-adaptation}
\Xi_1&:=&\sum_{k:\: 2^{-k}\leq h^*}\left(\frac{\| \K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{2^{-k}}\right)^{\frac{r-q_k}{2}}\int\left(
\overline{\Delta}_{\K,f}\big(2^{1-k} ,y\big)\right)^{q_k}\dd y
\nonumber\\
&=&\sum_{k:\: 2^{-k}\leq h^*}\int\left(
\overline{\Delta}_{\K,f}\big(2^{1-k} ,y\big)\right)^{r}\dd y
\leq c_7 \big(\|\K\|_\infty L\big)^{r}(h^*)^{\beta r -(r/p)+1}.
\end{eqnarray}
Applying Lemma \ref{lem:l_p-norm-of-bais}
with $\mathfrak{p}=r$, $s=\beta$ and $\cQ=L$ we get
\begin{eqnarray}
\label{eq8:proof-theorem-global-adaptation}
\Xi_2&:=&\sum_{k:\: 2^{-k}> h^*}\left(\frac{\| \K \|^2_{\infty}\eps^{2}
\ln (1/\eps)}{2^{-k}}\right)^{\frac{r-q_k}{2}}\int\left(
\overline{\Delta}_{\K,f}\big(2^{1-k} ,y\big)\right)^{q_k}\dd y
\nonumber\\
&=&c_8L^p\big(\|\K\|_\infty\big)^{r}\left(\eps^{2}
\ln (1/\eps)\right)^{\frac{r-p}{2}}\sum_{k:\: 2^{-k}> h^*}2^{-k\big[\beta p-\frac{r-p}{2}\big]}
\nonumber\\
&\leq&c_9L^p\big(\|\K\|_\infty\big)^{r}\left(\eps^{2}
\ln (1/\eps)\right)^{\frac{r-p}{2}}(h^*)^{\beta p-\frac{r-p}{2}}.
\end{eqnarray}
Here we have used that $\beta p-\frac{r-p}{2}<0.$
In view of (\ref{eq7:proof-theorem-global-adaptation}) and (\ref{eq8:proof-theorem-global-adaptation}) we choose $h^*$ from the equality:
$$
L^{r}(h^*)^{\beta r -(r/p)+1}=L^p\left(\eps^{2}
\ln (1/\eps)\right)^{\frac{r-p}{2}}(h^*)^{\beta p-\frac{r-p}{2}}.
$$
It yields $h^*=\left(L^{-1}\eps\sqrt{\ln (1/\eps)}\right)^{\frac{1}{\beta-1/p+1/2}}$ and we obtain finally that
\begin{eqnarray}
\label{eq9:proof-theorem-global-adaptation}
\Xi\leq c_{10}\big(\|\K\|_\infty\big)^{r}L^{\frac{r(1/2-1/r)}{\beta-1/p+1/2}}
\left(\eps\sqrt{\ln (1/\eps)}\right)^{\frac{r(\beta-1/p+1/r)}{\beta-1/p+1/2}}.
\end{eqnarray}
The assertion of the theorem in the case $(2\beta+1)p<r$
follows now from Theorem \ref{th:global-oracle-inequality},
(\ref{eq2:proof-theorem-global-adaptation}) and (\ref{eq9:proof-theorem-global-adaptation}). \epr
\section{Appendix}\label{sec:appendix}
\subsection{Proof of Lemma \ref{lem:bounds for bias}}
\paragraph{Proof of the first assertion}
The symmetry of the kernel $\K$ (Assumption~\ref{ass:assumption-on-kernel}~({\it 1})) implies
$$
S_{(-\ta^{\circ}, h)(\nu, h)}(\cdot)\equiv S_{(\ta^{\circ}, h)(-\nu, h)}(\cdot),\quad S_{(-\nu, h)}(\cdot)\equiv S_{(\nu, h)}(\cdot).
$$
Therefore it suffices to prove the first assertion of the lemma under the condition $\nu^\T\ta^{\circ}\geq 0$. In this case
$E_{(\ta^{\circ}, h)(\nu, h)}=\overline{E}_{(\ta^{\circ}, h)(\nu, h)}$ and we note that
\begin{equation}
\label{eq1:matrix for pseudo-estimator}
\overline{E}_{(\ta^{\circ}, h)(\nu, h)}
=
\left[E^{-1}_{(\ta^{\circ}, h)} + E^{-1}_{(\nu, h)}\right]^{-1}.
\end{equation}
For any $\theta=(\theta_1,\theta_2)\in\SS$ let $\ort{\ta}=(-\theta_2,\theta_1)$.
Using (\ref{eq1:matrix for pseudo-estimator}) we obtain
\begin{eqnarray*}
S_{(\ta^{\circ}, h)(\nu, h)}(x)
=
\int K(u)
f\big(h [\ta^{\circ} + \nu]^{\T} \ta^{\circ} u_1 + [\ort{\ta^{\circ}} + \ort{\nu}]^{\T} \ta^{\circ} u_2
+ x^{\T} \ta^{\circ}\big) \dd u
\\
=
\int\int \K(u_1)\K(u_2)f\big( h [1 + \nu^{\T} \ta^{\circ}] u_1 + \ort{\nu}^{\T} \ta^{\circ} u_2+ x^{\T} \ta^{\circ}\big) \dd u_1\dd u_2.
\end{eqnarray*}
We also have
$$
S_{(\nu, h)}(x)=
\int\int \K(u_1)\K(u_2)f\big( h \nu^{\T} \ta^{\circ} u_1 + \ort{\nu}^{\T} \ta^{\circ} u_2+ x^{\T} \ta^{\circ}\big) \dd u_1\dd u_2.
$$
Put
$
S_\nu^*(x)=\int \K(u_2)f\big(\ort{\nu}^{\T} \ta^{\circ} u_2+ x^{\T} \ta^{\circ}\big) \dd u_2
$
and consider two cases.
\medskip
$1^{0}.\;\; \ort{\nu}^{\T} \ta^{\circ}=0$. In this case $S_\nu^*(x)=f(x^{\T} \ta^{\circ})$ and
\begin{gather*}
S_{(\nu, h)}(x)=\int \K(u_1)f\big( h u_1 + x^{\T} \ta^{\circ}\big) \dd u_1=h^{-1}\int \K\big([t-x^{\T} \ta^{\circ}]/h\big)f(t) \dd t,
\\
S_{(\ta^{\circ}, h)(\nu, h)}(x)=\int \K(u_1)f\big( 2h u_1 +x^{\T} \ta^{\circ}\big) \dd u_1=(2h)^{-1}\int \K\big([t-x^{\T} \ta^{\circ}]/2h\big)f(t) \dd t.
\end{gather*}
Here we have used that $\ort{\nu}^{\T} \ta^{\circ}=0$ together with $\nu^{\T} \ta^{\circ}\geq 0$ implies $\nu=\ta^{\circ}$.
Thus, we obtain
\begin{eqnarray}
\label{eq2:proof-lemma-bais}
\left|S_{(\ta^{\circ}, h)(\nu, h)}(x)-S_{(\nu, h)}(x)\right|
&\le& \left|S_{(\ta^{\circ}, h)(\nu, h)}(x)-S_\nu^*(x)\right|+\left|S_{(\nu, h)}(x)-S_\nu^*(x)\right|
\nonumber\\*[2mm]
&\le&\Delta_{\K,f}\big(h,x^{\T} \ta^{\circ}\big)+\Delta_{\K,f}\big(2h,x^{\T} \ta^{\circ}\big)\leq 2\Delta^*_{\K,f}\big(2h,x^{\T} \ta^{\circ}\big).
\end{eqnarray}
\smallskip
$2^{0}.\;\; \ort{\nu}^{\T} \ta^{\circ}\neq0$. In this case we have
\begin{eqnarray*}
&&S_\nu^*(x)
=\int\int\frac{1}{h(1+\nu^{\T} \ta^{\circ})}\K\left(\frac{v_1}{h(1+\nu^{\T} \ta^{\circ})}\right)
\frac{1}{|\ort{\nu}^{\T} \ta^{\circ}|}\K\left(\frac{v_2-x^{\T} \ta^{\circ}}{|\ort{\nu}^{\T} \ta^{\circ}|}\right)f(v_2) \dd v_1\dd v_2,
\\*[2mm]
&&S_{(\vartheta^*, h)(\nu, h)}(x)
=\int \int \frac{1}{h(1+\nu^{\T} \ta^{\circ})}\K\left(\frac{v_1}{h(1+\nu^{\T} \ta^{\circ})}\right)
\frac{1}{|\ort{\nu}^{\T} \ta^{\circ}|}\K\left(\frac{v_2-x^{\T} \ta^{\circ}}{|\ort{\nu}^{\T} \ta^{\circ}|}\right)f(v_1+v_2) \dd v_1\dd v_2.
\end{eqnarray*}
Here we have used once again the symmetry of kernel $\K$. Thus, taking into account that $|\nu^{\T} \ta^{\circ}|\leq 1$, we get
\begin{eqnarray*}
&&\left|S_{(\ta^{\circ}, h)(\nu, h)}(x)-S_\nu^*(x)\right|
\\
&&\leq\int \frac{1}{|\ort{\nu}^{\T} \ta^{\circ}|}\left|\K\left(\frac{v_2-x^{\T} \ta^{\circ}}{|\ort{\nu}^{\T} \ta^{\circ}|}\right)\right|
\sup_{\delta\leq 2h}\left|\int\frac{1}{\delta}\K\left(\frac{v_1}{\delta}\right)
\big[f(v_1+v_2)-f(v_2)\big] \dd v_1\right|\dd v_2
\\*[2mm]
&&\leq\|\K\|_\infty\sup_{a>0}\left[\frac{1}{a}\int_{x^{\T} \ta^{\circ}-a/2}^{x^{\T} \ta^{\circ}+a/2} \sup_{\delta\leq 2h}\left|\int\frac{1}{\delta}\K\left(\frac{v_1}{\delta}\right)
\big[f(v_1+v_2)-f(v_2)\big] \dd v_1\right|\dd v_2\right].
\end{eqnarray*}
Here we have used that $\text{supp}(\K)\subseteq [-1/2,1/2]$ (Assumption \ref{ass:assumption-on-kernel} ({\it 1})). Hence,
\begin{eqnarray}
\label{eq3:proof-lemma-bais}
&&\left|S_{(\ta^{\circ}, h)(\nu, h)}(x)-S_\nu^*(x)\right|
\le \|\K\|_\infty\overline{\Delta}_{\K,f}\big(2h,x^{\T} \ta^{\circ}\big)\leq
\|\K\|_\infty\Delta^*_{\K,f}\big(2h,x^{\T} \ta^{\circ}\big).
\end{eqnarray}
If $\nu^{\T} \ta^{\circ}\neq 0$ we obtain by the same computations
$$
\left|S_{(\nu, h)}(x)-S_\nu^*(x)\right|\leq
\|\K\|_\infty\Delta^*_{\K,f}\big(h,x^{\T} \ta^{\circ}\big).
$$
Noting that $S_{(\nu, h)}(\cdot)\equiv S_\nu^*(\cdot)$ if $\nu^{\T} \ta^{\circ}= 0$ we get
\begin{eqnarray}
\label{eq33:proof-lemma-bais}
\left|S_{(\nu, h)}(x)-S_\nu^*(x)\right|\leq \|\K\|_\infty\Delta^*_{\K,f}\big(h,x^{\T} \ta^{\circ}\big),
\end{eqnarray}
that yields together with (\ref{eq3:proof-lemma-bais})
\begin{eqnarray}
\label{eq4:proof-lemma-bais}
\left|S_{(\ta^{\circ}, h)(\nu, h)}(x)-S_{(\nu, h)}(x)\right|\leq
2\|\K\|_\infty\Delta^*_{\K,f}\big(2h,x^{\T} \ta^{\circ}\big).
\end{eqnarray}
Finally, taking into account that in view of Assumption \ref{ass:assumption-on-kernel} ({\it 1}) $\|\K\|_\infty\geq 1$, we obtain
from (\ref{eq2:proof-lemma-bais}) and (\ref{eq4:proof-lemma-bais}) that
$$
\left|S_{(\ta^{\circ}, h)(\nu, h)}(x)-S_{(\nu, h)}(x)\right|\leq
2\|\K\|_\infty\Delta^*_{\K,f}\big(2h,x^{\T} \ta^{\circ}\big)\leq 2\|\K\|_\infty\Delta^*_{\K,f}\big(h^*_f,x^{\T} \ta^{\circ}\big),
$$
since we consider $h$ such that $2h\leq h^*_{f}$. The definition of $h^*_{f}$ implies
$$
\Delta^*_{\K,f}\big(h^*_f,x^{\T} \ta^{\circ}\big)\leq (h^*_f)^{-1/2} \| \K \|_{\infty}\eps \sqrt{\ln (1/\eps)}
$$
and the first assertion of the lemma follows.
\paragraph{Proof of the second and third assertions} In view of (\ref{eq33:proof-lemma-bais}) for $\forall\eta\leq h\leq h^*_{f}$
\begin{eqnarray*}
\left|S_{(\nu, \eta)}(x)-S_{(\nu, h)}(x)\right|&\leq&\left|S_{(\nu, \eta)}(x)-S_\nu^*(x)\right|+\left|S_{(\nu, h)}(x)-S_\nu^*(x)\right|
\\
&\le&\|\K\|_\infty\left[\Delta^*_{\K,f}\big(\eta,x^{\T} \ta^{\circ}\big)+\Delta^*_{\K,f}\big(h,x^{\T} \ta^{\circ}\big)\right]
\le
2\|\K\|_\infty\Delta^*_{\K,f}\big(h,x^{\T} \ta^{\circ}\big)
\\
&\leq& 2\|\K\|_\infty\Delta^*_{\K,f}\big(h^*_f,x^{\T} \ta^{\circ}\big)\leq 2(h^{*}_f)^{-1/2} \| \K \|^2_{\infty}\eps \sqrt{\ln (1/\eps)},
\end{eqnarray*}
in view of the definition of $ h^*_{f}$. The second assertion is proved.
\smallskip
We have for any $ h\leq h^*_{f}$
\begin{eqnarray*}
\big|S_{(\ta^{\circ}, h)} (x) - F(x)\big|
&=&
\left| \frac{1}{h}\int \K \left( \frac{u}{h}\right) \big[ f(u +x^{\T} \ta^{\circ}) - f(x^{\T} \ta^{\circ})\big] \dd u\right| \\%*[2mm]
&\le&\Delta_{\K,f}\big(h,x^{\T} \ta^{\circ}\big)
\le \Delta^*_{\K,f}\big(h,x^{\T} \ta^{\circ}\big)\leq \Delta^*_{\K,f}\big(h^*_f,x^{\T} \ta^{\circ}\big)
\\%*[2mm]
&=&(h^*_f)^{-1/2} \| \K \|_{\infty}\eps \sqrt{\ln (1/\eps)},
\end{eqnarray*}
in view of the definition of $ h^*_{f}$. The third assertion is proved. \epr
\subsection{Proof of Lemma \ref{lem:gauss-on-matrices}}
Since $\zeta_{y}(\cdot)$ is a zero mean Gaussian random function
we have
\begin{equation}
\label{eq1:proof-lemma-gauss-on-matrices}
\PP\left\{\sup_{E\in\mathcal{E}_{a,A}}\left|\zeta_{y}(E)\right|\geq u \right\}
\leq
2
\PP\left\{\sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E)\geq u \right\},\quad\forall u>0.
\end{equation}
By Lemma 12.2 in \cite{Lif} the median \( m \) of the random variable \( \sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E) \) is dominated by the expectation, that is \( m \le \EE \sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E) \). That along with the Borell, Tsirelson, Sudakov concentration inequality, see Theorem 12.2 in \cite{Lif} provides
\begin{equation}\label{eq02:proof-lemma-gauss-on-matrices}
\PP\left\{\sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E)\ge \EE \sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E) + z \right\}
\le \PP\left\{\sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E)\ge m + z \right\}
\le \PP\left\{ \varsigma \ge z\right \}
\end{equation}
since \( \sup_{E\in\mathcal{E}_{a,A}} \Var \left[ \zeta_{y}(E) \right] =1\). Here \( \varsigma\;\sim\;\norm{0}{1} \).
Thus, to complete the proof of the first assertion of the lemma it suffices to bound \( \EE \sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E) \). This will be done by the application of Dudley's theorem, see Theorem 14.1 in~\cite{Lif}.
Denote by \( \varrho \) the semi-metric generated by \( \zeta_{y}(\cdot) \) on \( \mathcal{E}_{a,A} \):
\begin{equation*}
\varrho(E,E^\prime)=\sqrt{\EE\left|\zeta_{y}(E)-\zeta_{y}(E^\prime)\right|^{2}}, \quad E,E^\prime\in\mathcal{E}_{a,A}.
\end{equation*}
Without loss of generality one can assume that \( \left| \det(E) \right| \ge \left|\det(E^{\prime})\right| \), then we have
\begin{eqnarray*}
&&\varrho^2(E,E^\prime)
=
2\left[1-\| \cc L\|_2^{-2}\sqrt{\left|\det(E)\right|\left|\det(E^{\prime})\right|}
\int \cc L(Ev) \cc L(E^{\prime}v)\dd v\right].
\\
&=&2\left[1-
\| \cc L\|_2^{-2}\sqrt{\frac{\left|\det(E^{\prime})\right|}{\left|\det(E)\right|}}
\int \cc L(z) \cc L(E^{\prime}E^{-1}z)\dd z\right]
\\
&=&2\left[1-\| \cc L\|_2^{-2}\sqrt{\frac{\left|\det(E^{\prime})\right|}{\left|\det(E)\right|}}
\int_{[-\frac{1}{2} , \frac{1}{2}]^{2}} \cc L(z) \cc L\big(E^{\prime}E^{-1}z\big)\dd z\right]
\\
&=&2\left[1- \frac{\sqrt{|\det (E^{\prime})|}}{\sqrt{|\det (E)|}} \right]
+
\frac{2}{\| \cc L \|^2_2} \frac{\sqrt{|\det (E^{\prime})|}}{\sqrt{|\det (E)|}}
\left[
\int_{[-\frac{1}{2} , \frac{1}{2}]^{2}} \cc{L}(z) \big[\cc{L}(z) - \cc{L}(E^{\prime}E^{-1}z)\big]\dd z
\right].
\end{eqnarray*}
One bounds the first summand with the use of the assumption \( |\det (E)| \ge a \):
\begin{equation*}
2\left[1- \frac{\sqrt{|\det (E^{\prime})|}}{\sqrt{|\det (E)|}} \right]
\le
\frac{2}{\sqrt{a} }\, \left|\sqrt{|\det (E)|} - \sqrt{|\det (E^{\prime})|}\right|
\le
\frac{2}{\sqrt{a} }\, \left|\det (E) - \det (E^{\prime})\right|^{1/2}.
\end{equation*}
As for the second term, putting
$$
\mathfrak{d}^2(E,E^\prime)
=
\int_{[-\frac{1}{2} , \frac{1}{2}]^{2}} \left| \cc L \big(E^{\prime}E^{-1}z\big)- \cc L(z)\right|^{2}\dd z,
$$
by the Cauchy-Schwartz inequality we get
$$
\int_{[-\frac{1}{2} , \frac{1}{2}]^{2}} \cc{L}(z) \left[ \cc L (z) - \cc L (E^{\prime}E^{-1}z) \right] \dd z
\le
\| \cc L \|_2 \mathfrak{d}(E,E^\prime).
$$
As $\|\cc L\|_2\geq 1$, we have
$$
\varrho^2(E,E^\prime)\leq 2a^{-1/2}
\left| \det (E)- \det (E^{\prime}) \right|^{1/2} + 2\mathfrak{d}(E,E^\prime).
$$
First, we note that
$$
\left|\det (E) -\det (E^{\prime})\right|\leq 4A\big|E-E^\prime\big|_\infty.
$$
Second, because \( \cc L \) satisfies the Lipschitz condition with a constant \( \Upsilon \), we have
$$
\mathfrak{d}(E,E^\prime)
\leq
\Upsilon \sup_{z\in[-\frac{1}{2} , \frac{1}{2}]^{2}}\left|(E^\prime-E)E^{-1}z\right|_2
\leq
2 \sqrt{2} \Upsilon Aa^{-1} \big|E-E^\prime\big|_\infty.
$$
Since we assumed $a\leq A$, the following bound holds:
\begin{equation}
\label{eq2:proof-lemma-gauss-on-matrices}
\varrho^2(E,E^\prime)
\leq
4 \big(\sqrt{2}\Upsilon+1\big)Aa^{-1}
\left(\big| E-E^\prime\big|^{1/2}_\infty \bigvee \big|E-E^\prime\big|_\infty \right).
\end{equation}
Consider the cube $[0,A]^{4}$ endowed with the vector supremum norm \( |z|_\infty=\max_{i=1,\ldots, 4}|z_i| \). Let
\( \mathfrak{E}_{[0,A]^{4},|\cdot|_\infty}(\cdot) \) denote the metric entropy of \( [0,A]^{4} \) measured in \( |\cdot|_\infty \). Then
\begin{equation*}
\mathfrak{E}_{[0,A]^{4},|\cdot|_\infty}(\epsilon)\leq 4\ln(A)+\left[ 4\ln{(1/(2\epsilon))} \right]_{+},\quad \forall\epsilon\in (0,1].
\end{equation*}
Denoting by \( \mathfrak{E}_{\mathcal{E}_{a,A},\varrho}(\cdot) \) the metric entropy of
\( \mathcal{E}_{a,A} \) measured in \( \varrho \), we get in view of (\ref{eq2:proof-lemma-gauss-on-matrices})
\begin{equation*}
\mathfrak{E}_{\mathcal{E}_{a,A},\varrho}(\delta)
\leq
\mathfrak{E}_{[0,A]^{4},|\cdot|_\infty}
\left(
\frac{\delta^{4}}{16(1 + \sqrt{2} \Upsilon)^2A^2a^{-2}}
\right),\quad\forall\delta\in (0,1],
\end{equation*}
and, therefore,
\begin{equation*}
\mathfrak{E}_{\mathcal{E}_{a,A},\varrho}(\delta)
\leq
4 \left[ \ln\big( A\vee\{A/a\}^2 \big) + \ln 8 + 2\ln{(1+\sqrt{2}\Upsilon)} +4\ln{(1/\delta)}\right].
\end{equation*}
Since \( \sup_{E\in\mathcal{E}_{a,A}} \Var \left[ \zeta_{y}(E) \right] =1\) the use of Dudley's integral bound, see Theorem 14.1 in~\cite{Lif}, leads to
\begin{eqnarray}
\label{eq100:proof-lemma-gauss-on-matrices}
\EE\Big[\sup_{E\in\mathcal{E}_{a,A}}\zeta_{y}(E)\Big]
&\le&
4\sqrt{2}\int_{0}^{1/2}\sqrt{\mathfrak{E}_{\mathcal{E}_{a,A},\varrho}(\delta)}\;\dd\delta
\nonumber\\
&\le&
4\sqrt{2} \left[ \ln(A\vee\{A/a\}^2) + 2 \ln{(1+ \sqrt{2} \Upsilon)} \right]^{1/2} + 29
\; =: \, \mathbf{c}(a,A).
\end{eqnarray}
Here we have used that \( \int_{0}^{1/2}\sqrt{\ln{(1/\delta)}}\;\dd\delta\leq 2^{-1}\sqrt{\pi} \).
The first assertion of the lemma follows now from \eqref{eq1:proof-lemma-gauss-on-matrices}, \eqref{eq02:proof-lemma-gauss-on-matrices}, \eqref{eq100:proof-lemma-gauss-on-matrices} and the standard bound for the Gaussian tail.
\smallskip
To justify the second assertion we first note that for any \( q\ge1 \)
$$
\mathbb{E}\Big[\sup_{E\in\mathcal{E}_{a,A}}\big|\zeta_{y}(E)\big|\Big]^{q}=
q\int_{0}^{\infty}u^{q-1}\mathbb{P}\left(\sup_{E\in\mathcal{E}_{a,A}}\big|\zeta_{y}(E)\big|\geq u\right)\dd u.
$$
Hence, applying the first assertion of the lemma we have
\begin{eqnarray*}
\mathbb{E}\Big[\sup_{E\in\mathcal{E}_{a,A}}\big|\zeta_{y}(E)\big|\Big]^{q}
&\le& \big[\mathbf{c}(a,A)\big]^{q} + q \int_0^{\infty}\PP\left\{ |\varsigma| \ge z\right \}(\mathbf{c}(a,A) + z)^{q-1} \dd z\\
&=& \mathbb{E}\big(\mathbf{c}(a,A)+|\varsigma|\big)^{q},
\end{eqnarray*}
where \( \varsigma\;\sim\;\norm{0}{1} \). Thus, we finally have
\begin{equation*}
\left(\mathbb{E}\Big[\sup_{E\in\mathcal{E}_{a,A}}\big|\zeta_{y}(E)\big|\Big]^{q}\right)^{1/q}
\le \mathfrak{c}_q \mathbf{c}(a,A).
\end{equation*} \epr
\subsection{Proof of Lemma \ref{lem:l_p-norm-of-bais}}
First, in view of the $(\mathfrak{p},\mathfrak{p})$-strong maximal inequality, see e.g. Theorem 9.16 in \cite{WheedenZygmund1977}, one has
\begin{equation*}
\left\|\overline{\Delta}_{\K,g}(h,\cdot)\right\|_\mathfrak{p}\leq\tau_\mathfrak{p}\left\|\Delta_{\K,g}(h,\cdot)\right\|_\mathfrak{p},
\end{equation*}
where the constant \( \tau_\mathfrak{p} \) depends only of \( \mathfrak{p} \).
Since $\Delta^*_{\K,g}(h,\cdot)\leq \overline{\Delta}_{\K,g}(h,\cdot)+\Delta_{\K,g}(h,\cdot)$ we have
\begin{equation}
\label{eq1:proof-of-lemma-l_p}
\left\|\Delta^*_{\K,g}(h,\cdot)\right\|_\mathfrak{p}\leq(\tau_\mathfrak{p}+1)\left\|\Delta_{\K,g}(h,\cdot)\right\|_\mathfrak{p}.
\end{equation}
For any $\delta\in (0,h]$ put
$
B(z,\delta)=\Big|\delta^{-1}\int\K\big([u-z]/\delta\big)\big(g(u)-g(z)\big)\dd u\Big|
$
and define
$$
\Delta^{(n)}_{\K,g}(h,z)=\sup_{\delta\in [hn^{-1},h]}B(z,\delta), \;\;n = 1,2, \ldots \;.
$$
We remark that the sequence \( \{\Delta^{(n)}_{\K,g}(h,\cdot)\}_{n\geq 1} \) increases monotonically and
\( \Delta^{(n)}_{\K,g}(h,z)\to\Delta_{\K,g}(h,z) \) for any \( z\in\RR \), as \( n\to\infty \).
Hence, by Beppo-Levi's theorem
$$
\left\|\Delta_{\K,g}(h,\cdot)\right\|_\mathfrak{p}=\lim_{n\to\infty}\left\|\Delta^{(n)}_{\K,g}(h,\cdot)\right\|_\mathfrak{p},
$$
and, in view of (\ref{eq1:proof-of-lemma-l_p}), to complete the argument we need to show that
\begin{equation}
\label{eq111:proof-of-lemma-l_p}
\sup_{g\in\mathbb{N}_\mathfrak{p}(s,\cQ)}\left\|\Delta^{(n)}_{\K,g}(h,\cdot)\right\|_\mathfrak{p}\leq 2\cQ h^{s}\|\K\|_\infty
\left[2^{s \mathfrak{p}}-1\right]^{-\frac{1}{\mathfrak{p}}},\;\;\forall n\geq 1.
\end{equation}
Assumption \ref{ass:assumption-on-kernel} ({\it 2}) implies that
we can assert that $B(z,\cdot)$ is continuous on $[n^{-1}h,h]$.
Hence for any $z\in \RR$ there exists $\delta(z)\in [n^{-1}h,h]$ such that
\begin{equation}
\label{eq2:proof-of-lemma-l_p}
\Delta^{(n)}_{\K,g}(h,z)=B\big(z,\delta(z)\big).
\end{equation}
For any \( l = 0, \ldots, \log_2n-1 \) (w.l.g. $\log_2n$ is assumed an integer) we consider the slices $V_l=\big\{z\in\RR:\;\; a_{l+1}<\delta(z)\leq a_l\big\}$ with \( a_l=2^{-l}h \). Later on the integration over empty set is supposed to be zero. Then
\begin{equation}
\label{eq22:proof-of-lemma-l_p}
\left\| \Delta^{(n)}_{\K,g}(h,\cdot) \right\|^{\mathfrak{p}}_\mathfrak{p}
=
\sum_{l=0}^{\log_2n-1} \int_{V_l} |B\big(z,\delta(z)\big)|^{\mathfrak{p}}\dd z.
\end{equation}
We will treat the cases $s\leq 1$ and $s>1$ separately. If $s<1$ on any slice \( V_l \), \( l = 0, \ldots, \log_2n \),
\begin{eqnarray}
\label{eq33:proof-of-lemma-l_p}
B\big(z,\delta(z)\big)
&\le&
\frac{\| \cc K \|_\infty}{\delta(z)}
\int_{-\frac{\delta(z)}{2} }^{\frac{\delta(z)}{2}}
\left|g( z + v )-g(z)\right|\dd v
\leq
\frac{2\| \cc K \|_\infty}{a_l} \int_{-\frac{a_l}{2} }^{\frac{a_l}{2}}
\left|g( z + v )-g(z)\right|\dd v \nn
&=& 2\|\K\|_\infty \int_{-\frac{1}{2}}^{\frac{1}{2}} \left|g( z + ta_l )-g(z)\right|\dd t.
\end{eqnarray}
We obtain from
(\ref{eq22:proof-of-lemma-l_p}) and (\ref{eq33:proof-of-lemma-l_p}) with the use of Minkowski's inequality for integrals and writing for ease of notation
$\mu=2\|\K\|_\infty$ that
\begin{eqnarray*}
\left\|\Delta^{(n)}_{\K,g}(h,\cdot)\right\|^{\mathfrak{p}}_\mathfrak{p}
&\le&
\mu ^{\mathfrak{p}}
\sum_{l=0}^{\log_2n-1}
\int
\left|
\int_{-\frac{1}{2}}^{\frac{1}{2}} \left| g (ta_l+z )-g(z) \right| \dd t
\right|^{\mathfrak{p}}\dd z
\nonumber
\\
&\le&
\mu^{\mathfrak{p}}
\sum_{l=0}^{\log_2n-1}
\left(
\int_{-\frac{1}{2}}^{\frac{1}{2}} \left\| g( \cdot + ta_l ) - g(\cdot) \right\|_\mathfrak{p} \dd t
\right)^{\mathfrak{p}}
\le \left[\frac{\cQ h^{s }\|\K\|_\infty 2^{1-s}}{(s+1)}\right]^\mathfrak{p}
\sum_{l=0}^\infty 2^{-ls \mathfrak{p}}.
\end{eqnarray*}
Here we have used that $g\in\mathbb{N}_\mathfrak{p}(s,\cQ)$.
Thus, we have for any $s\leq 1$ and any $n\geq 1$
\begin{equation}
\label{eq44:proof-of-lemma-l_p}
\sup_{g\in\mathbb{N}_\mathfrak{p}(s,\cQ)}
\left\|\Delta^{(n)}_{\K,g}(h,\cdot)\right\|_\mathfrak{p}\leq
2\cQ h^{s}\|\K\|_\infty \left[2^{s \mathfrak{p}}-1\right]^{-\frac{1}{\mathfrak{p}}}.
\end{equation}
If $s>1$, using Taylor's formula we have for any $g\in\mathbb{N}_\mathfrak{p}(s,\cQ)$ any $v\in\RR$
\begin{eqnarray*}
g(v+z)-g(z)
=\sum_{m=1}^{m_s}
\frac{g^{(m)}(z) }{m!}v^m
+
\frac{v^{m_s}}{(m_s-1)!}
\int_0^1 (1-\lambda)^{m_s -1} \left[g^{(m_s)}(z+v\lambda)-g^{(m_s)}(z)\right]\dd\lambda.
\end{eqnarray*}
We have in view of Assumptions \ref{ass:assumption-on-kernel} and \ref{ass2:assumption-on-kernel} for any $z\in\RR$
\begin{equation*}
B\big(z,\delta(z)\big)\leq\frac{\|\K\|_\infty}{(m_s-1)!}\frac{1}{\delta(z)} \int_{-\frac{\delta(z)}{2}}^{\frac{\delta(z)}{2}}
\int_{0}^{1} |v|^{m_s}(1-\lambda)^{m_s -1}\left|g^{(m_s)}\big(z+\lambda v\big)-g^{(m_s)}(z)\right|\dd\lambda\dd v.
\end{equation*}
By the latter inequality for any $z\in V_l$ we get
\begin{equation}
\label{eq3:proof-of-lemma-l_p}
B\big(z,\delta(z)\big)\leq \frac{2\|\K\|_\infty a^{m_s}_l }{(m_s-1)!}
\int_{-\frac{1}{2}}^{\frac{1}{2}}
\int_{0}^{1}|t|^{m_s}(1-\lambda)^{m_s-1}\left|g^{(m_s)}(z + \lambda ta_l )-g^{(m_s)}(z)\right|\dd\lambda\dd t.
\end{equation}
Thus, we obtain from (\ref{eq2:proof-of-lemma-l_p}), (\ref{eq22:proof-of-lemma-l_p}) and (\ref{eq3:proof-of-lemma-l_p}) with the use of Minkowskii inequality for integrals and denoting $\mu=2\|\K\|_\infty \big/(m_s-1)!$ that
\begin{eqnarray*}
&& \left\|\Delta^{(n)}_{\K,f}(h,\cdot)\right\|^{\mathfrak{p}}_\mathfrak{p}
=
\sum_{l=0}^{\log_2n-1} \int_{V_l} |B\big(z,\delta(z)\big)|^{\mathfrak{p}}\dd z
\\
&&\leq
\mu^{\mathfrak{p}}
\sum_{l=0}^{\log_2n-1} a^{m_s \mathfrak{p}}_l
\int
\left(
\int_{-\frac{1}{2}}^{\frac{1}{2}}
\int_{0}^{1}|t|^{m_s}(1-\lambda)^{m_s-1} \left|g^{(m_s)}(z + \lambda ta_l )-g^{(m_s)}(z)\right|\dd\lambda
\dd t\right)^{\mathfrak{p}}
\dd z
\nonumber \\
&&\leq
\mu^{\mathfrak{p}}
\sum_{l=0}^{\log_2n-1} a^{m_s \mathfrak{p}}_l
\left(
\int_{-\frac{1}{2}}^{\frac{1}{2}}
\int_{0}^{1}|t|^{m_s}(1-\lambda)^{m_s-1}\left\|g^{(m_s)}(\cdot + \lambda ta_l )-g^{(m_s)}(\cdot)\right\|_\mathfrak{p}\dd\lambda
\dd t
\right)^{\mathfrak{p}}
\nonumber\\
&&\leq \left[\frac{\cQ h^{s }\|\K\|_\infty 2^{1-s}}{(s+1)(m_s+1)(m_s-1)!}\right]^\mathfrak{p}
\sum_{l=0}^\infty 2^{-ls \mathfrak{p}} .
\end{eqnarray*}
Here we have used that $g\in\mathbb{N}_\mathfrak{p}(s,\cQ)$.
Thus, we have for any $s>1$ and $n\geq 1$
\begin{equation}
\label{eq4:proof-of-lemma-l_p}
\sup_{g\in\mathbb{N}_\mathfrak{p}(s,\cQ)
}\left\|\Delta^{(n)}_{\K,g}(h,\cdot)\right\|_\mathfrak{p}\leq 2 \cQ h^{s}
\|\K\|_\infty \left[2^{s \mathfrak{p}}-1\right]^{-\frac{1}{\mathfrak{p}}}.
\end{equation}
We conclude that (\ref{eq111:proof-of-lemma-l_p}) is established in view (\ref{eq44:proof-of-lemma-l_p}) and (\ref{eq4:proof-of-lemma-l_p}).
\epr
\bibliographystyle{agsm}
|
\section*{Outline}
This paper is concerned with mathematics of diffraction. More specifically we are interested
in the famous inverse problem for diffraction: given something that is putatively the diffraction
of something, what are all the somethings that could have produced this diffraction.
Diffraction has been a mainstay in crystallography for almost a hundred years.\footnote{The Knipping-Laue experiment establishing x-ray diffraction was
carried out in 1912.} With the proliferation of extraordinary new materials with varying degrees of order and disorder, the importance of diffraction in revealing internal structure continues to be central. The precision, complexity, and variety of modern diffraction images is striking, see for instance the recent review article \cite{Withers}. Nonetheless, in spite of many advances, the fundamental question of diffraction, the inverse problem of deducing physical structure from diffraction images, remains as challenging as ever.
The discovery of aperiodic tilings and quasicrystals revived interest in the mathematics of diffraction, particularly since the types of diffraction images generated by such structures -- essentially or exactly pure point diffraction together with symmetries not occurring in ordinary crystals --
had not been foreseen by mathematicians, crystallographers, or materials research scientists.\footnote{Indeed, the discovery of quasicrystals by Shechtman via diffraction experiments met with substantial disbelief for a period of about two years.
His determination to communicate it to a skeptical scientific community is particularly stressed by the committee awarding him the Nobel prize in Chemistry in 2011 for this discovery. }
Starting with \cite{Hof1},
the resulting
new techniques created in order to understand mathematical diffraction have been considerable.
Among them we mention the introduction of ideas from the harmonic analysis of translation bounded measures \cite{BM,BL}, ergodic dynamical systems \cite{Bell,Radin0,Radin,RW}, and stochastic point processes \cite{Gouere,XR, BBM}.
The diffraction of a structure is defined as the Fourier transform of its autocorrelation. The latter is a positive-definite measure in the ambient space of the structure -- typically $\mathbb{R}^3$, but mathematically any locally compact Abelian
group $\mathbb{G}$ -- and the diffraction is then a centrally symmetric positive measure in the corresponding Fourier reciprocal space -- either $\widehat{\mathbb{R}^3} \simeq \mathbb{R}^3$ or generally $\widehat{\mathbb{G}}$ as the case may be.
In this paper we examine three basic questions that arise from this theory. The first question
asks whether {\em every} positive centrally symmetric measure is actually the diffraction of something. The second asks about the classification of the all various
`somethings' that have this given diffraction. The third asks what kinds of structures
the `somethings' can be. In the case of pure point diffraction (when the diffraction measure
is a pure point measure) we can give a satisfactory answer to these problems.
We will be in the setting of locally compact
Abelian groups. The difficulties of the questions, at least as we examine them here, have little to do with the generality of the setting. They are just as hard for $\mathbb{R}^3$. Indeed the second question is a very old
one in crystallography, called the {\em homometry problem} and is part of {\em the} fundamental problem of diffraction theory, namely how to unravel the nature of a structure that has created a given diffraction pattern.
Our first question immediately raises the third question. Exactly what do we mean by a structure that can diffract? Typically diffractive structures are conceived of as discrete measures, for instance point measures describing the positions of atoms, vertices in tilings, etc., weighted by appropriate scattering strengths, or as continuous measures describing the (scattering) distribution of the material structure in space. More recently, starting with the work of Gou\'{e}r\'{e} \cite{Gouere0, Gouere} there has been a
shift to discussing the diffraction of certain point processes, which in effect are random point measures, and various distortions of these (see e.g. \cite{BBM,XR,LM}). In this paper we are led to introduce new structures which we call a {\em spatial stationary processes}. We shall show that these always lead
to autocorrelation and diffraction measures and contain the typical theories of diffraction
discussed above\footnote{There is a technical requirement about the existence
of second moments.}. In terms of these we get an attractive answer our three questions, in the case of pure point diffraction. However, the answer to the third question is interesting, because it does not describe the underlying structure explicitly, but rather in terms of the ways that it appears to compactly supported continuous functions on the ambient space. In this respect it resembles the description of measures given by the Riesz representation theorem or the theory of distributions with their spaces of test functions.
In our case, the convergence conditions of measures do not seem to be generally valid. In this respect there is much to be learned about the full scope of which
structures can diffract.
Let $\mathbb{G}$ be a locally compact Abelian group and $\widehat{\mathbb{G}}$ its dual group.
We treat these groups in additive notation. We begin with an abstract notion
of a stationary process for $\mathbb{G}$ and then shift our attention to the more concrete notion of
a spatial stationary process for $\mathbb{G}$ (\S\ref{Stochasticprocesses} and \S\ref{spatial}). As for terminology let us note that outside of the discussion of usual stochastic processes, the word `stochastic' is not used in our discussion of processes and stationary processes.
A spatial stationary process basically consists of a $\mathbb{G}$-invariant mapping
\[ N: C_c(\mathbb{G}) \longrightarrow L^2(X,\mu) \]
where $(X,\mu)$ is a probability space on which there is a measure preserving
action of $\mathbb{G}$.
The idea here derives directly from the theory
of stochastic point processes, and it is perhaps useful to discuss this briefly.
We follow \cite{Karr} here. A stochastic point
process, say in $\mathbb{R}^d$, is a random variable $M$ on a probability space
$(\Omega, \mathcal{F}, P)$ with values in the space $\mathcal{M}_p$ of point measures
$\xi$ on $\mathbb{R}^d$ for which $\xi(B) \in \mathbb{Z}_{\ge 0}$ for all bounded subsets $B$ of $\mathbb{R}^d$.
These are point measures and can be written in the form $\xi = \sum_{j=1}^K \delta_{x_j}$
where $K$ can be a positive integer or infinity and the points $x_j$ are (not necessarily distinct)
points of $\mathbb{R}^d$. Usually the space $\mathcal{M}_p$ is far larger than
needed. For one thing it allows points in $\mathbb{R}^d$ to appear with multiplicities and for another
it does not require any minimal separation between distinct points (a hard core condition) that
would usually be imposed in crystallography or in tiling theories. We assume then that we are
really only interested in some subset
$X \subset \mathcal{M}_p$ of permissible outcomes for the point process, and that almost surely
the point process produces measures in $X$. Typically we wish $X$ to be invariant under translation
since the position of our point sets is not particularly relevant, and we assume that the law $\mu$
of the process (the probability measure $\mu$ induced on $X$ by the random variable)
is invariant under translation (i.e. the process is stationary).
Now the point is this. If $f\in C_c(\mathbb{R}^d)$ then we obtain a random variable on $(\Omega, \mathcal{F}, P)$
by $\omega \mapsto N_f(\xi):= \sum_{j=1}^K f(x_j)$, where $\xi = M(\omega) \in X$ is the measure indicated above.
In effect $N$ can be viewed as a mapping from $C_c(\mathbb{R}^d)$ into mappings on $X$. Under
mild assumptions $N:C_c(\mathbb{R}^d) \longrightarrow L^2(\mathbb{R}^d,\mu)$. As is often the case with
random variables, we ignore the underlying probability space $(\Omega, \mathcal{F}, P)$
since we have everything we need from the probability space $(X,\mu)$ and the mapping $N$.
Then $N$ itself can be referred to as the stochastic point process.
In this paper $\mathbb{R}^d$ is replaced by any locally compact Abelian group $\mathbb{G}$.
Also $X$ is not assumed to have any particular
form (e.g. a set of point sets, or a set of measures). We do not wish to assume
any particular mathematical structure on the outcomes of the processes we are studying. Our point
of view is that we will start with a measure in $\widehat{\mathbb{G}}$ which we would like to show
is the diffraction of something. But we do not know in advance what sort of mathematical
objects could produce this diffraction (point sets, continuous distributions of density, Schwartz
distributions, etc.). We derive information
about $X$ only from the behaviour of the functions $N(F)$, $F\in C_c(\mathbb{G})$, upon it.
This is, physically, rather appropriate. In a physical situation one determines
the structure of a physical object by measurements upon it and the results of measurements
are ultimately all that one can know about it.
We are particularly interested in pure point diffraction. However, the concept of stationary processes
seems interesting in its own right and we develop the theory of these a bit, independent of the question of pure point diffraction. It is even useful initially to ignore the fact that the resulting process is {\it{spatial}},
that is, it is supposed to result from some structure in $\mathbb{G}$.
So at the beginning we often only require that we have a $\mathbb{G}$-invariant mapping $N:C_c(\mathbb{G}) \longrightarrow
\mathcal{H}$ where $\mathcal{H}$ is a Hilbert space with conjugation -- we call these
stationary processes. Assuming existence of second moments, we can then, for each such process, find an `autocorrelation measure' on $\mathbb{G}$ and a
`diffraction measure' on $\widehat{\mathbb{G}}$. The latter is always a centrally symmetric positive
measure, and through it we can define continuous and pure point diffraction. We show how pure point and continuous diffraction are tied
to notions of almost periodicity of $N$. Also we show how to decompose a sum of spatial processes into sub-processes which are respectively pure point and continuous.
Coming to the pure point part of the paper, \S\ref{Purepoint}, we assume given a centrally symmetric positive pure point measure $\omega$ on $\widehat{\mathbb{G}}$, which can be expressed as a Fourier transform.
We wish to create a probability space $(X,\mu)$ on which there is an ergodic action of $\mathbb{G}$ and a continuous stationary process $N: C_c(\mathbb{G}) \longrightarrow L^2(X,\mu)$ whose diffraction is $\omega$.
Not surprisingly phase factors play a crucial role in this since
it is phase information that is lost in the process of diffraction. There is a natural combinatorial type of group $Z$
generated by the positions of the Bragg peaks. This group can be seen as an abeleanized homotopy group of an associated Cayley graph.
It is the dual group of $Z$, or the dual group of a natural factor $\mathcal{Z}$ of this group,
that classifies all the spatial processes on $\mathbb{G}$ with diffraction $\omega$, the
classification being up to isomorphism or up to translational isomorphism respectively.
We call these phase factors {\em phase forms}. Background on phase forms is given in Section \ref{Cycles}. In \S\ref{CyclestoStochasticprocesses} we show how to
use $\omega$ and a phase form $a^*$ to construct a spatial process whose diffraction
is $\omega$. Uniqueness is already shown in \S\ref{Uniqueness}.
How all these ingredients give a solution to the homometry problem is discussed in
\S\ref{Homometry}.
Our solution to the pure point problem is in some sense an elaboration of the famous
Halmos-von Neumann result about realizing ergodic pure point dynamical systems in terms of the character theory of compact Abelian groups. There the objective was to
classify pure pointedness at the level of the spectrum. In our case we wish to classify the
diffraction, which amounts to not only knowing the spectrum but also the
intensity of diffraction at each point of the spectrum. Our approach to the construction
of a spatial process is to realize $X$ as the dual to the discrete group generated
by the set of Bragg peaks, but at the same time to include all the phase information which derives from the Bragg intensities. In this context we mention the recent work of Robinson \cite{Rob3} on how to realize systems with a given group of eigenvalues via so called cut-and-project schemes.
In \S\ref{compactG} we take a closer look at the situation that the underlying group is compact. In this case all processes can be realized as measure processes
under some weak assumption. We then specialize even further and in \S\ref{simplePeriodic}
discuss some results of Gr\"unbaum and Moore
on rational diffraction from one dimensional periodic structures. These pertain to the simple case that $\mathbb{G} = \mathbb{Z}/N\mathbb{Z}$, but even so, the results are interesting and not easy to obtain.
Finally in \S\ref{altConstruction} we sketch out a second approach to the construction of a spatial
process, this time based on Gel'fand theory.
Summarizing:
\begin{itemize}
\item Each pure point ergodic spatial process gives rise to a pure point diffraction measure $\omega$ and a phase form $a^*$.
\item Spatial processes with the same pure point diffraction measure and the same phase form are naturally isomorphic up to translation.
\item The set of all possible phase forms associated with a given pure point measure $\omega$ form an Abelian group $\mathcal{Z}(\omega)$, and for each choice of
$a^* \in \mathcal{Z}(\omega)$, there is a ergodic spatial process with diffraction $\omega$ and
phase form $a^*$.
\end{itemize}
The paper has two main features, after the introduction of stationary and spatial
processes: one is a development of a general theory of these types of processes
and the other is a detailed study of pure point diffraction and examples of how it
looks in special cases. Readers interested primarily in the pure point theory can skip
sections $3.3, 3.4 ,4,5$ on first reading if they so wish.
\section{Notation} \label{notation}
In this section we gather some notation used throughout.
\bigskip
$\mathbb{N}:= \{1,2, \dots, \}$. Let $\mathbb{G}$ be a locally compact abelian group.
We let $C_c(\mathbb{G})$ denote the space of compactly supported complex valued continuous
functions on $\mathbb{G}$. For compact $K \subset \mathbb{G}$, $C_K(\mathbb{G})$
is the subspace of elements of $C_c(\mathbb{G})$ whose support is in $K$. For $F\in C_c(\mathbb{G})$, $\widetilde F \in C_c(\mathbb{G})$ is defined
by $\widetilde F(x) = \overline{F(-x)}$. The translation action of $\mathbb{G}$ on itself is
denoted by $\tau$: $\tau(t)(x) = t +x$ for all $t,x \in \mathbb{G}$. This action determines, in the usual way, an action on $C_c(\mathbb{G})$.
By $l_\mathbb{G}$ we denote the (fixed) Haar measure on $\mathbb{G}$.
The convolution of two functions $F,G\in C_c (\mathbb{G})$ is defined to be the function $F\ast G\in C_c (\mathbb{G})$ given by
$$F\ast G (t) = \int_\mathbb{G} F( t- s ) G (s) d l_\mathbb{G} (s).$$
The dual group of $\mathbb{G}$ is denoted by $\widehat{\GG}$ and the \textit{Fouriert ransform} of an $F\in C_c (\mathbb{G})$ is denoted by $\widehat{F}$ i.e.
$$\widehat{F} (\gamma) = \int_\mathbb{G} \overline{(\gamma,t)} F(t) d l_\mathbb{G} (t).$$
A measure $\nu$ on $\mathbb{G}$ is called \textit{transformable} if there exists a measure $\sigma$ on $\widehat{\GG}$ with
$$ \int_{\widehat{\GG}} |\widehat{F}|^2 d\sigma = \int_\mathbb{G} F\ast \widetilde{F} d\nu$$
for all $F\in C_c (\mathbb{G})$.
In this case $\sigma$ is uniquely determined and called the \textit{Fourier transform of $\nu$}. In such a situation we call $\sigma$ \textit{backward transformable}. Unlike all other notation introduced in this section the concept of backward transformability is not standard. However, it is clearly very useful in our context as it is exactly the property shared by the diffraction measures we investigate.
A sequence $\{A_n\}$ of compact subsets
of $\mathbb{G}$ is called a \textit{van Hove sequence}
if for every compact $K\subset \mathbb{G}$
\begin{equation} \label{van-hove}
\lim_{n\to\infty} \frac{l^{}_\mathbb{G}
(\partial^K A_n)}
{l^{}_\mathbb{G} (A_n)} \; = \; 0.
\end{equation}
Here,
for compact $A,K$, the
``$K$-boundary'' $\partial^K A $ of $A$ is defined as
$$ \partial^K A := ((A + K)\setminus A^\circ) \cup
(\overline{\mathbb{G}\setminus A} - K) \cap A.$$
The existence of van Hove
sequences for all $\sigma$-compact locally compact Abelian groups is shown in
\cite[p.~249]{Martin}, see also Section 3.3 and Theorem (3.L)
of \cite[Appendix]{Tempelman}.
\section{Stationary processes introduced}\label{Stochasticprocesses}
In this section we introduce our concept of stationary process on the locally compact Abelian group $\mathbb{G}$. To do so we first discuss some background on unitary representations.
The concept of stationary process is somewhat more general than needed in the remainder of the paper. The reason for this is twofold: On the one hand this general treatment does not lead to more complicated proofs but rather makes proofs more transparent. On the other hand for a the study of mixed spectra this general concept may prove to be especially useful.
\bigskip
Let $\mathbb{G}$ be a locally compact abelian group and let $\mathcal{H}$ be a Hilbert space with
inner product $\langle\cdot | \cdot \rangle$. We assume linearity in the first variable
and conjugate linearity in the second.
A strongly continuous unitary representation of the group $\mathbb{G}$ on $\mathcal{H}$ is a map $T$ from $\mathbb{G}$ into the bounded operators on $\mathcal{H}$ such that $r \mapsto \langle T_r f, T_r f\rangle$ is constant on $\mathbb{G}$ for any $f\in \mathcal{H}$, $T_0$ is the identity on $\mathcal{H}$, $T_{t+s} = T_t T_s$ holds for all $s,t\in \mathbb{G}$ and $t\mapsto T_t f$ is continuous for any $f\in \mathcal{H}$. Then, obviously each $T_s$ is unitary (i.e. bijective and norm preserving).
An $f\in \mathcal{H}$ is then called an {\it eigenfunction\/} of $T$ with
{\it eigenvalue\/} $\xi\in \widehat{\GG}$ if $T_t f = \overline{(\xi,t)} f$ for every
$t\in \mathbb{G}$. The closed subspace of $\mathcal{H}$ spanned by all eigenfunctions is denoted by $\mathcal{H}_p$. The representation $T$ has {\it pure point spectrum} if $\mathcal{H}= \mathcal{H}_p$, or equivalently if $\mathcal{H}$ has an orthonormal basis consisting of eigenfunctions \footnote{Often the eigenfunction condition is written
as $T_t f = (\xi,t) f$ for eigenvalue $\xi\in \widehat{\GG}$. For our purposes things work
out more smoothly by formulating it with the conjugate value.}.
By Stone's theorem, compare \cite[Sec.~36D]{Loomis}, there exists a projection-valued measure
\begin{equation*}
E_T\! : \; \mbox{Borel sets of $\widehat{\GG}$} \; \longrightarrow \;
\mbox{projections on $\mathcal{H}$}
\end{equation*}
with
\begin{equation} \label{spectralmeasure}
\langle f \mid T_t f \rangle \; = \;
\int_{\widehat{\GG}} (\xi, t) \,{\rm d} \rho^{}_f (\xi)\,
\end{equation}
for all $t\in\mathbb{G}$ and $f\in \mathcal{H}$,
where $\rho^{}_f$ is the measure on $\widehat{\GG}$ defined by
$\rho^{}_f (B) := \langle f \mid E_T (B)f\rangle$. Then $\rho^{}_f$ is called the \textit{spectral measure} of $f$. It is the unique measure on $\widehat{\mathbb{G}}$ satisfying \eqref{spectralmeasure}.
A mapping $\overline{(\cdot)}$ of a Hilbert space $(\mathcal{H}, \langle \cdot \mid \cdot \rangle)$ into itself is said to be a {\it conjugation} of $\mathcal{H}$ if
\begin{itemize}
\item[{(i)}] $\overline{(\cdot)}$ is conjugate-linear and $\overline{\overline{\xi}} = \xi$;
\item[{(ii)}] for all $\xi, \psi \in \mathcal{H}$, $\langle \overline{\xi} \mid \overline{\psi}\rangle
= \langle \psi \mid \xi\rangle$.
\end{itemize}
\begin{definition} Let $\mathbb{G}$ be a locally compact abelian group. A {\it stationary process}
or $\mathcal{H}$-{\it stationary process} on $\mathbb{G}$ is a triple $\mathcal{N}=(N, \mathcal{H}, T)$ consisting of a Hilbert space $\mathcal{H}$ with
conjugation, a strongly continuous unitary representation $T$ of $\mathbb{G}$ on $\mathcal{H}$ and a linear
$\mathbb{G}$-map $N: C_c (\mathbb{G})\longrightarrow \mathcal{H}$ such that for all
$F\in C_c (\mathbb{G})$ $N(\overline F) = \overline{N(F)}$. A stationary process is called {\it ergodic} if the eigenspace of $T$ for the eigenvalue $0$ is one-dimensional.
\footnote{The origins of the definition lie in the theory of stochastic point processes,
hence the name. We will often simply write {\em process}
instead of {\em stationary process}. We do not use the word {\em stochastic}, although
our definition allows for processes that are stochastic in nature.}
\end{definition}
\medskip
The set of all $\mathcal{H}$-stationary processes on $\mathbb{G}$ forms a real vector space in the obvious
way by taking linear combinations of mappings. There is also a canonical notion of isomorphism between stochastic processes: The processes $(N_1, \mathcal{H}_1, T_1)$ and $(N_2,\mathcal{H}_2, T_2)$ are called \textit{isomorphic} if there exists an invertible unitary map $U : \mathcal{H}_1\longrightarrow \mathcal{H}_2$ which intertwines $T_ 1$ and $T_2$ and satisfies $N_2 = U N_1$.
\medskip
We will be interested in the spectral theory of stationary processes. An important notion is given next.
\begin{definition} A stationary process $\mathcal{N}= (N,\mathcal{H},T)$ is said to have {\it pure point spectrum} if the
representation $T$ of $\mathbb{G}$ on $\mathcal{H}$ has pure point spectrum.
\end{definition}
\bigskip
\begin{definition} A stationary process $\mathcal{N}=(N,\mathcal{H},T)$ is called a {\it spatial
stationary process} if there is a probability space $(X,\mu)$ and an action
$T: \mathbb{G} \times X \longrightarrow X, (t,x)\mapsto t\cdot x$ such that $\mu$ is invariant
under the $\mathbb{G}$-action on $X$ through $T$, $\mathcal{H} = L^2(X,\mu)$ with the usual inner product $\langle f \mid g\rangle = \int_X f \overline{g} d\mu$,
and the action of $T$ is extended to an action on $\mathcal{H}$ by $T_t f(\cdot) = f((-t)(\cdot))$ for all $t\in\mathbb{G}$, $f\in L^2(X,\mu)$. The conjugation
here is the natural one: complex conjugation of functions on $X$. We write
$\mathcal{N}=(N,X,\mu,T)$. The stationary
spatial process is called {\it full} if the algebra generated by functions of the
form $\psi \circ N(F)$, $\psi \in C_c(\mathbb{C})$, $F\in C_c(\mathbb{G})$, is dense in
$L^2(X,\mu)$.
\end{definition}
\begin{remark}
In some sense the assumption of fullness is only a convention. More precisely, given any spatial stationary process we may create a full process with essentially the same properties by a variant of Gelfand construction. Details are given below in Theorem~ \ref{fullProcess}.
\end{remark}
For the sake of economy, we have used the same notation, $T$, for the action of $\mathbb{G}$ on $X$
and the corresponding action of $\mathbb{G}$ on $\mathcal{H} := L^2(X,\mu)$. We often omit the word `stationary' in the sequel, but we will always understand
it to be in force.
The definition of stationary
processes implicitly defines them as {\em real} processes: $N(F)$ is real for all real-valued
functions $F$ on $\mathbb{G}$ in the sense that $\overline{N(F)} = N(F)$. In the case of
spatial processes, all real-valued functions on $\mathbb{G}$ are mapped by $N$ to real-valued functions on $X$. One could relax the conditions to allow complex-valued processes, but we shall not do that
here.
\bigskip
\begin{definition}\label{definition-isomorphism}
Two stationary spatial processes $\mathcal{N} = (N,X,\mu,T)$ and $\mathcal{N}'=(N',X',\mu',T')$ are {\it spatially isomorphic} if there is an invertible unitary mapping
$M : L^2 (X,\mu)\longrightarrow L^2 (X',\mu')$ with
\begin{itemize}
\item $M\circ N = N'$;
\item $M \circ T_t = T_t' \circ M$ for all $t\in \mathbb{G}$;
\item $M(f g) = M(f) M(g)$ for all $f,g\in L^\infty (X,\mu)$ and $M^{-1} (f' g') = M^{-1} (f') M^{-1} (g')$ for all $f',g'\in L^\infty (X',\mu')$.
\end{itemize}
The map $M$ is then called spatial isomorphism between the processes.
\end{definition}
The isomorphism class of $\mathcal{N}= (N,X,\mu,T)$ is denoted by $[\mathcal{N}]=[(N,X,\mu,T)]$.
\begin{remark} (a) The first two points of the definition just say that the processes are isomorphic. The special requirement for spatial isomorphy is the third point. As shown in the next proposition, the spatial isomorphism are exactly those isomorphisms which are compatible with the $L^\infty$ module structure of $L^2$ in the sense that $M(f g) = M(f) M(g)$ for all $f\in L^\infty$ and $g\in L^2$.
(b) Our definition of spatial isomorphism parallels the
notion of spectral isomorphism for measure preserving group actions
on probability spaces. Our conditions on $M$ imply that spatial isomorphism implies conjugacy in the sense
that there is a measure isomorphism between the Borel sets of $X$ and
those of $X'$ \cite{Walters}, Theorem~ 2.4. If in addition $X,X'$ are complete separable metric spaces then this can be improved to being an isomorphism between $X$ and $X'$,
that is a bijective intertwining mapping between two subsets of full measure
in $X$ and $X'$ respectively \cite{Walters}, Theorem~2.6.
\end{remark}
\begin{prop} Let $\mathcal{N} = (N,X,\mu,T)$ and $\mathcal{N}'=(N',X',\mu',T')$ be spatially isomorphic processes with spatial isomorphy $M$. Then, $M$ maps $L^{\infty} (X,\mu)$ into $L^\infty (X',\mu')$ and $M(f g) = M(f) M(g)$ holds for all $f\in L^{\infty} (X,\mu)$ and $g\in L^2 (X,\mu)$. The corresponding statements hold for $M^{-1}$ as well.
\end{prop}
\begin{proof} It suffices to consider the statements on $M$. By assumption we have that for $f,g\in L^\infty$ the product $M(f) M(g)$ belongs to $L^2$ and in fact equals $M (f g)$ (and similarly for $M^{-1}$). By a simple limit argument, we then see that for any $f\in L^\infty$ the equation $M (f g) = M(f) M(g)$ must hold for any $g\in L^2$. This shows that the operator of multiplication by $M(f)$ is defined on the whole of $L^2$ (as $M$ is onto). Furthermore, the graph of every multiplication operator can easily be seen to be closed. Thus, the closed graph theorem gives that the operator of multiplication by $M(f)$ is a bounded operator on $L^2$. This in turn yields that $M(f)$ belongs to $L^\infty$.
\end{proof}
\section{Diffraction theory for stochastic processes}
In this section we show how each stationary process with a second moment comes with a positive definite measure $\gamma$ on $\mathbb{G}$ called the autocorrelation, a positive measure $\omega$ on $\widehat{\mathbb{G}}$ called the diffraction measure, and the diffraction-to-dynamics map $\theta$. We then discuss how the autocorrelation comes about as a limit and characterize existence of a second moment.
\bigskip
\subsection{Autocorrelation and diffraction}
Let a stationary process $\mathcal{N}=(N,\mathcal{H},T)$ be given.
We say that $\mathcal{N}$ has a \textit{second moment} (or is with second moment) if there is a measure $\mu^{(2)}:=\mu^{(2)}_N$
on $C^{\mathbb{R}}_c(\mathbb{G} \times \mathbb{G})$ (necessarily unique if it exists) satisfying
\begin{equation}\label{2momentCondition}
\mu^{(2)}(F \otimes G) = \langle N(F) \mid N(G) \rangle \,
\end{equation}
for all real valued $F,G\in C_c (\mathbb{G})$.
Much of this paper is based on the assumption that the second moment measure exists.
Since $N$ is real, so is $\mu^{(2)}$, and hence it is real valued
\footnote{The $m$th moment of $\mathcal{N}$ will be defined by
$\mu^{(m)}(F_1\otimes \cdots \otimes F_m) = \int_X N(F_1) \cdots N(F_m) \, d \mu$
which is different from what we get from the inner product $\langle \cdot \mid \cdot \rangle$
when $m=2$ unless we restrict to real-valued functions. This is why we restrict to real-valued
functions in this definition.}. Furthermore $\mu^{(2)}$ is evidently
translation bounded since $N$ is $\mathbb{G}$-invariant.
\begin{prop} \label{def:gamma}
Let $\mathcal{N} = (N,\mathcal{H},T)$ be a process.
Then $\mathcal{N}$ has a second moment if and only if there is a measure $\gamma = \gamma_N$ on $\mathbb{G}$
satisfying
\begin{equation}\label{acEquation}
\gamma(F * \tilde G) = \langle N(F) \mid N (G)\rangle
\end{equation}
for all $F,G \in C_c(\mathbb{G})$.
In the case that such a measure exists, it is unique and furthermore it is real (i.e. assigns real values to real valued functions), positive definite, and translation bounded.
\end{prop}
\begin{proof} It suffices to restrict attention to real-valued $F$ and $G$ in $C_c (\mathbb{G})$ as $N$ is invariant under a conjugation and hence both $F\ast \tilde{G}$ and $\langle N(F) \mid N (G)\rangle $ are linear in $F$ and antilinear in $G$.
Let $\psi := \mu^{(2)}$ above.
The assumption of stationarity
of $N$ shows that $\psi$ is invariant under simultaneous translation of both of its two variables.
Define the mapping
\begin{eqnarray}
\mathbb{G} \times \mathbb{G} &\longrightarrow& \mathbb{G} \times \mathbb{G}\\
(x,y) &\mapsto & (x-y,y) =:(u,v) \nonumber
\end{eqnarray}
For $F_1,F_2 \in C_c^{\mathbb{R}}(\mathbb{G})$ we define $H \in C_c(\mathbb{G}\times \mathbb{G})$ by
$H(u,v) = H(x-y,y) = F_1(x)F_2(y)$. This change of variables defines a new measure
$\psi^*$ via
\[ \int_\mathbb{G} \int_\mathbb{G} F_1(x)F_2(y) \,{\rm d}\psi(x,y) = \int_\mathbb{G} \int_\mathbb{G} H(u,v) \,{\rm d} \psi^*(u,v) \,.\]
Since $(\tau_t x,\tau_t y) \leftrightarrow (u,\tau_t v)$, the translation invariance of $\psi$
translates to translation invariance of the second variable of $\psi^*$. Thus for
measurable sets $A,B \subset \mathbb{G}$ we have
$\psi^*(A \times B) = \psi^*(A \times (t+B))$ for all $t\in \mathbb{G}$. For $A$ fixed this is leads to a translational
invariant measure on $\mathbb{G}$ which is hence a multiple $c(A)$ of Haar measure
$l_\GG$:
\[ \psi^*(A \times B) = c(A) l_\GG(B) \,. \]
The mapping $A \mapsto c(A)$ is a measure on $\mathbb{G}$, and this measure
is the reduction $\psi^{\mbox{red}}$ of $\psi$:
\[\psi^*(A\times B) = \psi^{\mbox{red}}(A)\, l_\GG(B) \,.\]
Now for $H(u,v) = F_1(u+v)F_2(v)$,
\begin{eqnarray*}
\int\int H(u,v) \,{\rm d} \psi^*(u,v) &=& \int\int F_1(u+v) F_2(v) \,{\rm d} \psi^{\mbox{red}}(u) \,{\rm d} l_\GG(v)\\
&=&\int\left( \int F_1(u+v) F_2(v) \,{\rm d} l_\GG(v)\right) \,{\rm d} \psi^{\mbox{red}}(u)\\
&=& \int (F_1*\tilde{F_2})(u) \,{\rm d} \psi^{\mbox{red}}(u) \, ,
\end{eqnarray*}
so finally
\[ \psi(F_1 \otimes F_2) = \psi^{\mbox{red}}(F_1 *\tilde{F_2}) \,.\]
and $\psi^{\mbox{red}}$ is the desired measure, which we now
denote by $\gamma$.
It is positive definite since $\gamma(F * \tilde F) = \mu^{(2)}(F \otimes F)
= ||(N(F)||^2\ge 0$. All positive definite measures are translation bounded \cite{BF}.
Since $\psi := \mu^{(2)}$ is a real measure, one sees that $\gamma$ is also
a real measure. The denseness of the set of functions $F*\tilde G$ in $C_c(\mathbb{G})$ shows
that $\gamma$ is unique.
We can reverse all the steps of this proof, and from the existence of the real positive
definite measure $\gamma$ satisfying \eqref{acEquation} derive the corresponding
second moment measure \eqref{2momentCondition}.
\end{proof}
Existence of a second moment immediately implies some continuity properties of $N$.
\begin{coro}\label{continuity-N} Let $(N,\mathcal{H},T)$ be a stationary process with second moment. Let $1 \leq p,q \leq \infty$ be given with $1/p + 1/q = 1$. Then, for any compact $K\subset \mathbb{G}$, there exists a $C_K\geq 0$ with
$$ |\langle N(F) \mid N(G)\rangle |\leq C_K \|F\|_{L^p (\mathbb{G})} \| G\|_{L^q (\mathbb{G})}$$
for all $F,G\in C_c (\mathbb{G})$ with support contained in $K$. In particular the map $N : C_c (\mathbb{G})\longrightarrow L^2 (X,m)$ is continuous (with respect to the $L^2$-norm on $C_c(\mathbb{G})$).
Also $N:C_K(\mathbb{G}) \longrightarrow L^2 (X,m)$ is continuous with
respect to the sup norm on $C_K(\mathbb{G})$.
\end{coro}
\begin{proof} For $F,G\in C_c (\mathbb{G})$ with support contained in $K$ the support of $F\ast \widetilde{G}$ is contained in $K - K$ and $\|F\ast \widetilde{G}\|_\infty \leq \|F\|_{L^p (\mathbb{G})} \|G\|_{L^q (\mathbb{G})}$. As $|\langle N(F), N(G)\rangle | = |\gamma (F\ast \widetilde{G} ) |$, this easily gives the first statement. Now, the second statement follows by taking $p = q = 2$. Finally,
for $F\in C_K(\mathbb{G})$, $||F||_{L^2(\mathbb{G})} \le ||F||_\infty (l_{\mathbb{G}}(K))^{1/2}$.
\end{proof}
The continuity property of the preceding corollary allows one to extend the map $N$. This is discussed next.
\begin{coro} \label{extendingN}
Let $\mathcal{N}= (N,\mathcal{H},T)$ be a stationary process with second moment. The map $N$ can be uniquely extended to the vector space of all measurable functions $f$ on $\mathbb{G}$ all of which vanish outside some compact set and are square integrable with respect to the Haar measure
on $\mathbb{G}$.
This extension (again denoted by $N$) is $\mathbb{G}$-equivariant and satisfies
\[ \int_\mathbb{G} F * \tilde F \,{\rm d} \gamma = \langle N(F) | N(F) \rangle \, ,\]
meaning in particular that the integral exists and is finite. \end{coro}
\begin{proof} It suffices to show that $N$ can be uniquely extended to $L^2 (U, l_U)$ for any open $U$ in $\mathbb{G}$ with compact closure. Let such an $U$ be given.
Let $F\in L^2 (U, l_U)$ be given. As, $U$ is open, we can then find a sequence $\{F_n\}$ in $C_c (U)$ converging to $F$ in the sense of $L^2 (U, l_U)$.
By the previous corollary, we infer then that $\{N (F_n)\}$ must be a Cauchy-sequence, whose limit does not depend on the choice of the approximating sequence $\{F_n\}$. This shows that $N$ can be extended to $L^2 (U, l_U)$.
\smallskip
By construction of $N(F)$ and the definition of $\gamma$ we have furthermore
$$\langle N (F)\mid N(F)\rangle = \lim_{n\to \infty} \int_\mathbb{G} F_n \ast \tilde F_n d\gamma.$$
Moreover, a direct application of Cauchy-Schwartz inequality shows that $F_n \ast \tilde F_n$ converges to $F \ast \tilde F$ with respect to the supremum norm and has support contained in $\overline{U}- \overline{U}$. This easily gives that
$$\lim_{n\to \infty} \int_\mathbb{G} F_n \ast \tilde F_n d\gamma = \int_\mathbb{G} F\ast \tilde F d\gamma.$$
This finishes the proof.
\end{proof}
\begin{remark}
Note that any bounded measurable function with compact support belongs to $L^2 (U,d\,l_\mathbb{G})$ for any open $U$ containing the support. Thus, $N$ can in particular be extended to bounded measurable functions with compact support.
\end{remark}
The above discussion shows that any stationary process with a second moment gives rise to a real positive definite measure $\gamma$. As $\gamma$ is positive definite it is transformable i.e. its Fourier transform $\omega$ exists \cite{BF,SM}, and since $\gamma$ is a real measure, $\omega$
is centrally symmetric ($\omega(A) = \omega(-A)$
for all measurable sets $A$). The measures $\gamma$ and $\omega$ lie at the heart of our investigation.
\begin{definition} Let $(N,\mathcal{H},T)$ be a stationary process with second moment. Then, $\gamma:=\gamma_N$ is called the {\it autocorrelation} of the stationary process. Its Fourier transform $\omega:=\omega_N:=\widehat{\gamma_N}$ is called the {\it diffraction} or {\it diffraction measure} of the stationary process.
\end{definition}
\begin{definition} A stationary process $(N,\mathcal{H},T)$ with second moment is said to have {\it pure point diffraction}
(resp. {\it continuous diffraction}) if the diffraction measure $\omega$ is a pure point measure
(resp. continuous measure).
\end{definition}
The definition of $\omega$ yields that it is a centrally symmetric positive measure and is the unique measure satisfying
\begin{equation}\label{omega-gamma}
\int_{\widehat{\mathbb{G}}} |\widehat F|^2 \,{\rm d}\omega = \int_{\mathbb{G}} F * \widetilde F \,{\rm d} \gamma \, ,
\end{equation}
or equivalently, by linearizing this,
\begin{equation*}
\int_{\widehat{\mathbb{G}}} \widehat F \overline{\widehat G} \,{\rm d} \omega
= \int_\mathbb{G} F*\tilde G \,{\rm d}\gamma
\end{equation*}
for all $F,G \in C_c(\mathbb{G})$. Considering the Hilbert space $L^2(\widehat{\mathbb{G}}, \omega)$ with the corresponding inner product
being written $\langle \cdot \mid \cdot \rangle_{\widehat{\mathbb{G}}}$ and using the definition of $\gamma$, we obtain
for any $F,G \in C_c(\mathbb{G})$,
\begin{equation} \label{innerProduct}
\langle \widehat F \mid \widehat G \rangle_{\widehat{\mathbb{G}}}
= \int_{\widehat{\mathbb{G}}} \widehat F \overline{\widehat G} \,{\rm d} \omega\\
= \int_\mathbb{G} F*\tilde G \,{\rm d}\gamma\\
= \langle N(F) \mid N(G) \rangle \,.
\end{equation}
This shows that the map $\widehat{F} \mapsto N(F)$ is an isometry.
We define an action $\hat T$ of $\mathbb{G}$ on $L^2(\widehat{\mathbb{G}},\omega)$ by defining
$\hat T_t f$ for all $f\in L^2(\widehat{\mathbb{G}},\omega)$ through
\[\hat T_t f(k) = \overline{\langle k, t \rangle} f(k) \]
for all $t \in \mathbb{G}$ and all $k\in \widehat{\mathbb{G}}$.
\begin{remark} We note that if $N$ is a $\mathcal{H}$-stationary process on $\mathbb{G}$
with second moment
then for all $c\in \mathbb{R}$, $cN$ is another such process. The corresponding
autocorrelation and diffraction measures are then scaled by $|c|^2$.
\end{remark}
\subsection{The diffraction to dynamics map}\label{diff2dyn}
Any process with a second moment comes with an isometric $\mathbb{G}$-map $\theta$ from $L^2 (\widehat{\mathbb{G}},\omega)$ to $\mathcal{H}$. This allows one to transfer certain questions from $\mathcal{H}$ to $L^2 (\widehat{\mathbb{G}},\omega)$. Also, it means that for certain eigenvalues there are canonical eigenfunctions available. These topics are studied next.
\begin{prop} \label{diffractiontodynamics} Let $\mathcal{N}=(N,\mathcal{H},T)$ be a stationary process with
second moment.
Then $\{\widehat{F}: F\in C_c (\mathbb{G})\}$ is dense in $L^2 (\widehat{\mathbb{G}},\omega)$ and
there exists a unique isometric embedding
\begin{eqnarray}\label{difmToDyn}
\theta=\theta_N : L^2(\widehat{\mathbb{G}}, \omega)\longrightarrow \mathcal{H}
\end{eqnarray}
with $\widehat{F}\mapsto N(F)$ for each $F\in C_c (\mathbb{G})$.
This mapping is a $\mathbb{G}$-map and furthermore, for all $D\in L^2(\widehat{\mathbb{G}},\omega)$,
\[\theta(\widetilde{D}) = \overline{\theta(D)}\,.\]
\end{prop}
\begin{proof} We begin with the denseness statement.
Let $L$ be the closure of $\{\widehat{F}: F\in C_c (\mathbb{G})\}$ in $L^2 (\widehat{\mathbb{G}},\omega)$.
We wish to show that $L=L^2(\widehat{\mathbb{G}},\omega)$.
As $\omega$ is a translation bounded measure, the space $C_c (\widehat{\mathbb{G}})$ is dense in $L^2 (\widehat{\mathbb{G}},\omega)$. Thus it suffices to show that $C_c (\widehat{\mathbb{G}})$ belongs to $L$.
As the convolutions of elements from $ C_c (\mathbb{G})$ belong to $C_c (\mathbb{G})$ as well, we infer that for all $G_1,\ldots, G_m \in C_c (\mathbb{G})$
$$ \widehat{F_1}\cdots \widehat{F_n} \cdot \widehat{G_1}\cdots \widehat{G_m} \in L
$$
for all $F_i\in C_c (\mathbb{G})$. Now $\widehat{F_j}$ belongs to the set $C_0 (\widehat{\mathbb{G}})$ of continuous functions on $\widehat{\mathbb{G}}$ vanishing at infinity. An application of the Stone-Weierstrass theorem then shows that
$$ f \widehat{G_1}\cdots \widehat{G_m}\in L$$
for each $f\in C_0 (\widehat{\mathbb{G}})$. Fixing $f\in C_c (\widehat{\mathbb{G}})$ we can then use another application of the Stone-Weierstrass theorem, to infer that
$$ f g\in L$$
for all $f\in C_c (\widehat{\mathbb{G}})$ and $g \in C_0 (\widehat{\mathbb{G}})$. This yields $C_c (\widehat{\mathbb{G}})\subset L$ as we intended to show.
The existence of $\theta$ now follows by defining it initially
by $\theta: \widehat F \mapsto N(F)$ for all $F \in C_c(\mathbb{G})$,
using the fact that it is an isometric mapping by \eqref{innerProduct}, and then extending it
to the $L^2$-closure $L$ which is $L^2(\widehat{\mathbb{G}},\omega)$.
As for the second statement, first we note that the $\mathbb{G}$-action on $L^2(\widehat{\mathbb{G}}, \omega)$ is
designed to make the mapping $\hat F \mapsto F$ into a $\mathbb{G}$-map, while
$F\mapsto N(F)$ is already a $\mathbb{G}$-map, and second, for all $F\in C_c(\mathbb{G})$,
$\widetilde{\widehat{F}} = \widehat{\overline{F}}$, from which
\[\theta\left(\widetilde{\widehat{F}}\right) = \theta( \widehat{\overline{F}})
= N({\overline{F}}) = \overline{N(F)} = \overline{\theta(\widehat F)}\,.\]
We finish using the denseness of the first part.
\end{proof}
\begin{definition} The map $\theta=\theta_N$ associated to a stationary process $\mathcal{N}=(N,\mathcal{H},T)$ with second moment is called the {\it diffraction to dynamics map}.
\end{definition}
The relevance of $\theta$ for spectral theoretic considerations comes from the following consequence of the previous proposition.
\begin{theorem} \label{theta-as-spectralmeasure} Let $\mathcal{N}=(N,\mathcal{H},T)$ be a stationary process with
second moment and $\theta$ the associated diffraction to dynamics map. Then, for any $H\in L^2 (\widehat{\mathbb{G}},\omega)$ the spectral measure $\rho^{}_{\theta (H)}$ is given by
$$ \rho^{}_{\theta (H)} = |H|^2 \omega.$$
\end{theorem}
\begin{proof} As $\theta$ is a $\mathbb{G}$-map and an isomorphism a short calculation gives
$$ \langle \theta (H) \mid T_t \theta (H)\rangle = \langle \theta (H)\mid \theta ( (t,\cdot) H) \rangle = \int_{\widehat{\mathbb{G}}} (t,\xi) |H|^2 d \omega.$$
Thus, the (inverse) Fourier transform of $|H|^2 \omega$ is given by $t\mapsto \langle \theta (H) \mid T_t \theta (H)\rangle$.
By the characterization of the spectral measure in \eqref{spectralmeasure} the theorem follows.
\end{proof}
From the previous theorem it is clear that if $\mathcal{H}$ has pure
point spectrum then $\omega$ is discrete and the diffraction is pure point. However,
it is not clear that pure point diffraction implies pure point spectrum but as we shall
see, it is so: each implies the other in the case of spatial processes.
\medskip
The diffraction to dynamics map connects pure point diffraction to eigenvectors
of the $\mathbb{G}$ action on $\mathcal{H}$. Later, we shall see that in the case of
spatial processes, this allows us to make a complete correspondence between
pure point diffraction and pure point dynamics.
For a pure point diffractive stationary processes with diffraction measure $\omega$ we define
the set of its {\it atoms} by
$$\mathcal{S} =\mathcal{S}(\omega) :=\{k\in \widehat{\mathbb{G}}: \omega (\{k\}) \ne 0\} = \{k\in \widehat{\mathbb{G}}: \omega (\{k\}) > 0\} .$$
The set $\mathcal{S}$ is often called the {\it Bragg spectrum} of the process, and
its elements are sometimes called {\it Bragg peaks}\footnote{Sometimes, the term Bragg peak
is also taken to mean both the position $k$ of the atom and its intensity $\omega(\{k\})$.}.
In the sequel we will often omit the brackets when dealing with one element sets of the form $\{k\}$. In particular, we will set $ \omega(k):=\omega(\{k\})$
for $k\in \widehat{\mathbb{G}}$.
Let $1_k$ be the characteristic function of the set $\{k\} \subset \widehat{\mathbb{G}}$ i.e. $1_k(k')$ is
$1$ or $0$ according as $k'$ equals $k$ or not. It is easy to see that these functions
are the only possible eigenfunctions for our action of $\mathbb{G}$ on $L^2(\widehat{\mathbb{G}}, \omega)$.
Define $f_k$, $k\in \mathcal{S}$, by $f_k = \theta(1_{{k}})$.
Then each $f_k$ is an eigenfunction in
$\mathcal{H}$ for the eigenvalue $k$, in the sense that $T_t f_k = \overline{\langle k,t\rangle}f_k$.
\medskip
\begin{lemma}\label{fminusk} Let $\mathcal{N}=(N,\mathcal{H},T)$ be a stationary process with
second moment. Then,
for all $k \in \mathcal{S}$, $-k \in \mathcal{S}$,
and $f_{-k} = \overline{f_k}$ and $||f_{-k}|| = ||f_k|| =\omega(k)^{1/2}$.
\end{lemma}
\begin{proof} Since $\omega$ is centrally symmetric, $\omega(\{-k\}) = \omega(\{k\})$,
which gives the first statement. For the second, using Prop. \ref{diffractiontodynamics}
we have $f_{-k} = \theta(1_{-k}) = \theta(\widetilde{1_k}) =
\overline{\theta(1_k)} = \overline{f_k}$. Then
$||f_{-k}||^2 = \langle f_{-k} \mid f_{-k}\rangle = \langle f_{k} \mid f_{k}\rangle =
||f_k||^2 = \langle 1_k, 1_k\rangle = \omega(k)$.
\end{proof}
We can use the preceding considerations to compute the map $N$ in the case of pure point diffraction.
\begin{prop} \label{eigenExpansionOfN(F)}
Let $(N,\mathcal{H},T)$ be a stationary process with pure point diffraction and associated Bragg peaks $\mathcal{S}$. Let $\theta$ be the associated diffraction to dynamics map and $f_k = \theta (1_k)$, $k\in \mathcal{S}$. Then
$$ N(F)=\sum_{k\in \mathcal{S}} \widehat{F} (k) f_k$$
for all $F\in C_c (\mathbb{G})$.
\end{prop}
\begin{proof} By definition of $\theta$ we have $N(F) = \theta(\widehat{F})$. Obviously, $\widehat{F} = \sum_{k\in \mathcal{S}} \widehat{F} (k) 1_k$ and the claim follows.
\end{proof}
\subsection{Spatial stationary processes: Two-point correlation} \label{spatial}
In this section we show how the autocorrelation can be given a meaning that agrees with
`classical' two-point correlation associated to stationary point processes.
\bigskip
Assume that we are given an ergodic spatial stationary process $\mathcal{N}= (N,X,\mu,T)$.
As discussed in Corollary \ref{extendingN} we can assume that $N$ is defined on all measurable bounded functions with compact support.
We also assume that $\mathbb{G}$ has a countable basis of topology and is hence metrizable. Let $\{A_n\}_{n=1}^\infty$ be a van Hove sequence for $\mathbb{G}$ (as discussed in Section \ref{notation}). We assume that
this is fixed once and for all.
\begin{definition} \label{acDef} For $\xi \in X$ we define the {\em two-point correlation or autocorrelation} of $\xi$
at any $F \in C_c(\mathbb{G})$ as
\[\lim_{B \to 0} \frac{1}{l_\GG(B)} \left( \lim_{n\to\infty} \frac{1}{l_\GG(A_n)}
\int_{A_n}
(N(1_B) N(F))(T_{-x} \xi) \,{\rm d} l_\GG(x)\right) \,\]
whenever the limit exists.
\end{definition}
This needs some comments: The limits in the definition are taken in the order indicated: first the inner and then the outer.
$B$ is an open (or measurable) neighbourhood of $0$ in $\mathbb{G}$. The statement
$B\to 0$ means that we take a nested descending sequence $\{B_m\}$ of such neighbourhoods, all within some fixed compact set $K$, and that $l_\GG(B_m) \to 0$. We are using the notation $\,{\rm d} x$ to stand for the longer $\,{\rm d} l_\GG(x)$.
The definition requires that we give meaning to $N(1_B)$ as a measurable
function on $\mathbb{G}$. This uses the extension of $N$ from $C_c (\mathbb{G})$ to $L^2$-functions with compact support given in Corollary \ref{extendingN}.
The intuition behind the definition is as follows. The two-point correlation at
$\xi \in X$ for $F\in C_c(\mathbb{G})$
should look something like
\[ \lim_{n\to\infty} \frac{1}{l_\GG(A_n)} \int_{A_n}\int_{A_n} \xi(x)\xi(y) F(-x+y) \,{\rm d} x
\,{\rm d} y
= \lim_{n\to\infty} \frac{1}{l_\GG(A_n)} \int_{A_n} \xi(x)
\int_\mathbb{G} \xi(y) (T_x F)(y) \,{\rm d} y
\,{\rm d} x \, ,\]
where the right hand side arises by using the usual trick from van Hove sequences and the compactness of the support
of $F$. (That is, the difference of the two sides of the equation is due to the difference
between $-A_n$ and $-A_n+ \supp(F)$, which by the van Hove assumption is irrelevant in the limit.) Of course in our case $\xi$ is not a function of $x$ and the integrands
do not make sense. But the inner integral on the right-hand side is what should
be $N(T_x F)(\xi) = N(F)(T_{-x} \xi)$ and we can rewrite this `autocorrelation' as
\[\lim_{n\to\infty} \frac{1}{l_\GG(A_n)} \int_{A_n} (T_{-x}\xi)(0) N(F)(T_{-x} \xi)
\,{\rm d} l_\mathbb{G}(x) \, .\]
The term $(T_{-x}\xi)(0)$ has no meaning. But Palm theory tells
us how to go around this. We instead average over a small neighbourhood $B$
of $0$, and this brings us to Definition~\ref{acDef}.
\smallskip
\begin{theorem} \label{acThm} Let $\mathbb{G}$ be a locally compact group whose topology has a countable basis and let $(N,X,\mu,T)$ be an ergodic spatial stationary process on
$\mathbb{G}$ with second moment.
The two-point correlation of $\xi \in X$ exists $\mu$-almost surely and
its value at $F\in C_c(\mathbb{G})$ is $\gamma(F)$.
\end{theorem}
\begin{proof} The function
$(N(1_B)N(F))(T_{-x} \xi)$ is measurable as a function of $x\in \mathbb{G}$ and the Birkhoff
theorem says that almost surely
\begin{equation} \label{BET}
\lim_{n\to\infty}\frac{1}{l_\GG(A_n)} \int_{A_n} (N(1_B) N(F))(T_{-x} \xi) \,\,{\rm d} x
= \int_X N(1_B)N(F) \,{\rm d} \mu \, ,
\end{equation}
meaning that the limit will exist and equal the right hand side. We shall prove that
\begin{equation}\label{Bto0}
\lim_{B \to 0} \frac{1}{l_\GG(B)}\int_X N(1_B)N(F) \,{\rm d} \mu
= \gamma(F) \, ,
\end{equation}
proving that the definition of Definition~\ref{acDef} works almost
surely in $\xi$ for each $F \in C_c(\mathbb{G})$ and that $\gamma$ does have
the meaning of an autocorrelation.
This has to be made to work simultaneously for all $F$ in $C_c(\mathbb{G})$, which will be shown in the usual way from a countable basis of $C_c(\mathbb{G})$. This will prove
Theorem ~\ref{acThm}.
\smallskip
Here are the details: It is enough to prove \eqref{Bto0} for real-valued
$F$. Using \eqref{innerProduct} we have for all
$F,G\in C^\mathbb{R}_c(\mathbb{G})$,
\begin{equation*} \label{linearization}
\int_\mathbb{G} F*\tilde G \,{\rm d} \gamma = \langle N(F) \mid N(G) \rangle \,.
\end{equation*}
Let $F\in C_c^\mathbb{R}(\mathbb{G})$ and choose $\epsilon >0$. Let $B$ be a measurable subset of $\mathbb{G}$ with compact
closure. Then by Prop.~\ref{extendingN}, $N(B):= N(1_B)$ is defined and is
a measurable $L^2$-function on $X$. By linearization
we have
\begin{eqnarray*}
\int_X N(1_B)N(F) \,{\rm d} \mu &=& \int_\mathbb{G} 1_B *\tilde F \,{\rm d}\gamma
\nonumber\\
&=& \int_\mathbb{G} \int_\mathbb{G} 1_B(x)\tilde F(y-x) \,{\rm d} x \,{\rm d} \gamma(y)
=\int_\mathbb{G} \int_\mathbb{G} 1_B(x) F(x-y) \,{\rm d} x \,{\rm d} \gamma(y) \, .
\end{eqnarray*}
Since $F$ is uniformly continuous on $\mathbb{G}$, for any sufficiently small
neighbourhood $B$ of $0$, $|F(x-y) - F(-y)| <\epsilon$ for all $x\in B$ and
for all $y \in \mathbb{G}$. Then
\[
\left|\int_\mathbb{G} 1_B(x) F(x-y) \,{\rm d} x -l_\GG(B) F(-y) \right|
\le \int_\mathbb{G} 1_B(x) | F(x-y) - F(-y)| \,{\rm d} x < \epsilon\, l_\GG(B)\]
for all $y \in \mathbb{G}$, so
\begin{eqnarray*}
\frac{1}{l_\GG(B)}&&\left|\int_X N(B)N(F) \,{\rm d} \mu
- l_\GG(B) \int_\mathbb{G} F(-y) \,{\rm d}\gamma(y) \right| \\
&=&\frac{1}{l_\GG(B)} \left| \int_\mathbb{G} \int_\mathbb{G} 1_B(x) F(x-y) \,{\rm d} x \,{\rm d} \gamma(y) - l_\GG(B) \int_\mathbb{G} F(-y) \,{\rm d}\gamma(y) \right| \\
&\le& \frac{1}{l_\GG(B)} \int_\mathbb{G} \int_\mathbb{G} 1_B(x) | F(x-y) - F(-y)| \,{\rm d} x
\,{\rm d}\gamma(y)\\
& \le& \frac{1}{l_\GG(B)} \epsilon \,l_\GG(B)
|\gamma|(-\supp(F) +B).
\end{eqnarray*}
Since $\gamma(y) = \gamma(-y)$ ($\gamma$ is positve-definite and real, Prop.~\ref{def:gamma}),
we see that as $B \to 0$ (and correspondingly $\epsilon \to 0$)
we have a proof of \eqref{Bto0}. Together with \eqref{BET} we have
the desired interpretation of Def.~\ref{acDef} as the autocorrelation
at $\xi$ and its almost sure equality with $\gamma(F)$, and Prop.~\ref{acThm}
is proved.
\end{proof}
\begin{remark} This type of result was first established for certain uniformly discrete point processes on Euclidean space by Hof \cite{Hof1} (see \cite{Martin} for the case of general locally compact abelian groups). This was then extended to rather general point processes by Gou\'{e}r\'{e} \cite{Gouere} and to certain measure processes by Baake / Lenz \cite{BL}. A unified treatment was then given by Lenz / Strungaru \cite{LS}. Our result contains all these results (provided the underlying process is real).
\end{remark}
\subsection{A second glance at second moments}
The discussion above has shown that existence of second moments has strong consequences. It implies existence of
the diffraction to dynamics map by Proposition \ref{diffractiontodynamics} and further continuity properties
by Corollary \ref{continuity-N}. It turns out that a converse of sorts holds. This is investigated in this
section. Along the way, we will also show that continuity of $N$ implies an intertwining property of the map $N$ and this will be crucial for our considerations.
\medskip
Any process $(N, \mathcal{H}, T)$ gives rise to two representations of $C_c (\mathbb{G})$ (and in fact even of $L^1 (\mathbb{G})$): The representation $L$ lives on $L^2 (\mathbb{G})$ and acts by
$$ L_G F = G\ast F$$
for $F,G\in C_c (\mathbb{G})$.
The representation $T$ (extending the action $T$ and therefore denoted by the same letter) acts by
$$ T_G f := \int G (t) T_t f \, dl_\mathbb{G} (t) $$
for $G\in C_c (\mathbb{G})$ and $f\in \mathcal{H}$. Continuity of $N$ now yields an intertwining property:
\begin{lemma} \label{intertwiner} Let $(N, \mathcal{H}, T)$ be a process with a continuous $N$. Then,
$$N\circ L_G = T_G \circ N$$
for all $G\in C_c (\mathbb{G})$.
\end{lemma}
\begin{proof} We have
\[N(L_G (F)) = N(G \ast F) = N\left(\int_\mathbb{G} G(t)T_t F \,d\,l_\mathbb{G}(t)\right)\,.\]
By the linearity and continuity of $N$, it commutes with taking integrals and
\[N\left(\int_\mathbb{G} G(t)T_t F dl_\mathbb{G}(t)\right) = \int_\mathbb{G} G(t) N(T_t F)\, dl_\mathbb{G}(t) =\int_\mathbb{G} G (t) T_t N(F) dl_\mathbb{G} (t) \]
holds, which finishes the proof.
\end{proof}
We will need a somewhat stronger continuity property of $N$.
\begin{definition} Let $1\leq p,q\leq \infty$ with $1/p + 1/q = 1$ be given. The process $(N, \mathcal{H}, T)$ is said to be weakly $(p,q)$-continuous if for any
compact $K\subset \mathbb{G}$ there exists a $C_K$ with
$$|\langle N(F) \mid N(G)\rangle| \leq C_K \|F \|_{L^p (\mathbb{G}) } \|G\|_{L^q (\mathbb{G})}$$
for all $F,G\in C_c (\mathbb{G})$ with support contained in $K$ (see Cor.~\ref{continuity-N}).
\end{definition}
\begin{remark} (a) Note that continuity of $N$ is exactly weak $(2,2)$-continuity.
(b) By standard interpolation theory, we can conclude that weak $(2,2)$ continuity together with weak $(1,\infty)$ continuity implies
weak $(p,q)$ continuity for all $p,q$ with $1\leq p,q\leq \infty $ and $1/p + 1/q = 1$.
\end{remark}
\begin{theorem} Let $(N, \mathcal{H}, T)$ be a stochastic process. Then, the following assertions are equivalent:
\begin{itemize}
\item[(i)] $N$ has a second moment.
\item[(ii)] $N$ is weakly $(p,q)$-continuous for all $1\leq p,q\leq \infty$ with $1/p + 1/q = 1$.
\end{itemize}
In this case $N$ is continuous and has the intertwining property that
$$ N \circ L_G (F) = T_G N (F)$$
for all $F,G\in C_c (\mathbb{G})$.
\end{theorem}
\begin{proof} (i)$\Longrightarrow$ (ii): This follows immediately from Corollary \ref{continuity-N}.
\smallskip
(ii)$\Longrightarrow$ (i): As $N$ is weakly $(2,2)$ continuous, it is continuous and the intertwining property follows from the previous lemma.
In the remaining part of the proof we will only consider real valued functions. By Proposition \ref{def:gamma}, it suffices to show
existence of a measure $\gamma$ on $\mathbb{G}$ with
$$\langle N(F) \mid N(G)\rangle = \gamma (F\ast \tilde G)$$
for all $F,G\in C_c (\mathbb{G})$. A short calculation shows that for such a $\gamma$ the equality
$$ \langle N(F) \mid N(G)\rangle = \int F(y) (\gamma \ast G) (y) dl_\GG(y)$$
must hold for all $F,G\in C_c (\mathbb{G})$. The idea is now to `reverse' this reasoning to conclude existence of $\gamma$. The main issue
is to show continuity of the object $H_G$ replacing the not yet defined $\gamma \ast G$. This is shown using the intertwining property. Here are the details:
We choose an arbitrary compact $K$ in $G$ and assume without loss of generality that $0\in K$. Let $U$ be an open neighborhood of $0$ with compact closure and set
$K_1 $ to be the closure of $K + U$. Set $L = K_1 - K_1 + K_1 - K_1$.
As $N$ is weak $(1,\infty)$ continuous, and $L^\infty (L)$ is the dual of $L^1 (L)$, we can find to each $G\in C_c (\mathbb{G})$ with support in
$K_1$ a function $H_G$ in $L^\infty (L)$ with
$$ \langle N(F) \mid N(G)\rangle = \int F(y) H_G (y) dl_\GG (y)$$
for all $F\in C_c (\mathbb{G})$ with support in $L$ and
\begin{equation}\label{star2.9} \|H_G\|_\infty \leq C \|G\|_\infty, \end{equation}
where $C = C (L)$. As $N$ has the intertwining property, a short argument shows that
$$H_{G_1\ast G_2} (x) = \int G_1 (y) H_{G_2} (x - y) dl_\GG (y)$$
for $x\in K$ for all $G_2 \in C_c (\mathbb{G})$ with support contained in $K$ and all $G_1\in C_c (\mathbb{G})$ with support contained in $U$. This gives
in particular, that $H_{G_1\ast G_2}$ is a continuous function (on $K$) for all such $G_1, G_2$. Now
\eqref{star2.9} yields
$$\| (H_{D \ast G} - H_G)\|_\infty = \| H_{D\ast G- G} \|_\infty \leq C \| D \ast G - G\|_\infty.$$
Taking an approximate unit for $D$, we wee that $H_G$ can be approximated uniformly by continuous functions. This shows continuity of $H_G$. Moreover, by $\eqref{star2.9}$ again
the map
$$G \mapsto H_G (x)$$
is continuous for each $x$. Thus, we can indeed define the measure $\gamma$ with
$$\gamma \ast G (x) = H_G (x).$$
By construction $\gamma$ has the desired properties.
\end{proof}
\section{Almost periodicity}\label{Almost}
Almost periodicity is closely linked to nature of diffraction via the Fourier transform. This concept allows one to compare properties of the autocorrelation and the diffraction measure. In particular, it can be used to characterize pure point and continuous diffraction. The considerations of this section play a role in subsequent parts of the paper: They are used in Section \ref{pc:split} to decompose an arbitrary processes into a part with pure point diffraction and a part with continuous diffraction, and in Section \ref{Purepoint} to prove the equivalence of pure point diffraction spectrum and pure point spectrum. The material of this section derives from \cite{GA} and, in a more accessible account \cite{SM}, and from \cite{LR,LS} as well. For further studies of aspects of almost periodicity in our context we refer to \cite{Gouere,Sol,Lag}.
\bigskip
Almost periodicity is most commonly defined by a compactness condition: $f \in C_u(\mathbb{G})$ is almost periodic if the translation orbit $\mathbb{G}.f$ of $f$ is compact. Here $C_u(\mathbb{G})$
is the space of uniformly continuous $\mathbb{C}$-valued functions on $\mathbb{G}$. The key point,
however, is in which topology is the translation orbit to be compact? In the case of the
sup-norm topology the concept defines {\it strong almost periodicity}. In the case
that the topology is defined by the family of semi-norms induced by the set of
all continuous linear functionals on $C_u(\mathbb{G})$, it is called {\it weak almost periodicity}.
Strong almost periodicity coincides with H.~Bohr's original concept of almost periodicity:
$f\in C_u(\mathbb{G})$ is strongly periodic if and only if for every $\epsilon >0$ the set
of $t\in \mathbb{G}$ for which $||T_t f - f||_{\sup} < \epsilon$ is relatively dense. It is harder
to get an intuitive feel for weak almost periodicity. Fortunately what one needs to know
about it is fairly straightforward:
\begin{itemize}
\item all positive definite functions on $\mathbb{G}$ are weakly almost periodic;
\item for every weakly almost periodic function $f$ the {\it mean} of $f$ defined as
\[ \lim_{n\to\infty} \frac{1}{l_\GG(A_n)} \int_{A_n} f(x+t) \,{\rm d}l_\GG(t) \,\]
exists for any van Hove (more generally F\o lner) sequence $\{A_n\}$ in $\mathbb{G}$ and any $x\in \mathbb{G}$ and
its value is independent of both $x$ and the choice of van Hove sequence.
\end{itemize}
One says that a weakly almost periodic function $f$ is {\it null weakly almost periodic} if its mean is $0$. These concepts lift to measures in a simple way: a Borel measure $\phi$ on $\mathbb{G}$ is called {\it strongly} (resp. {\it weakly}, {\it null weakly}) {\it almost periodic} if $f*\phi$ is a strongly (resp. weakly, null weakly) almost periodic function
for every $f\in C_c(\mathbb{G})$.
\smallskip
Positive definite measures on $\mathbb{G}$ turn out to be
weakly almost periodic and are also, as we have noted already, Fourier transformable.
\begin{theorem} \label{thm-GA} {\rm (Gil de Lamadrid, Argabright \cite{GA})}
Every weakly almost periodic measure $\phi$ is uniquely expressible as the sum of a strongly
almost periodic measure and null weakly almost periodic measure. If the measure $\phi$ is
Fourier transformable then so too are the strong and null weak components, and
furthermore the Fourier transforms of these components are the pure point and continuous parts of $\widehat\phi$. \qed
\end{theorem}
These considerations can be applied to strongly continuous unitary representations and hence to processes as well. This is discussed next. The crucial connection is given by the following (well known) lemma.
\begin{lemma} \label{Fft-is-wap} Let $T$ be a strongly continuous unitary representation of $\mathbb{G}$ on $\mathcal{H}$ then for any $f \in \mathcal{H}$, the function $t\mapsto \langle f \mid T_t f\rangle =:F_f (t) $ is positive definite and hence weakly almost periodic and Fourier transformable. Its Fourier transform is the spectral measure $\rho^{}_f$.
\end{lemma}
\begin{proof} As $T$ is unitary, we have for any $t_1,\ldots, t_n\in \mathbb{G}$ and $c_1,\ldots, c_n\in \mathbb{C}$
$$ \sum_{k,l=1}^n c_k \overline{c_l} F_f (t_k - t_l) = \|\sum_{l=1}^n c_l F_f (t_l)\|^2 \geq 0$$
and $F_f$ is shown to be positive definite. The remaining statements follow from the above discussion.
\end{proof}
Combining the previous lemma and the previous theorem, we infer the following corollary.
\begin{coro} Let $T$ be a strongly continuous unitary representation of $\mathbb{G}$ on $\mathcal{H}$ then for any $f \in \mathcal{H}$, the function $t\mapsto \langle f \mid T_t f\rangle$ is the sum of a strongly almost periodic function and a null weakly almost periodic function. This strongly almost periodic function is given by
the (inverse) Fourier transform of the pure point part of the spectral measure $\rho^{}_f$ and the null weakly almost periodic function is given by the (inverse) Fourier transform of the continuous part of the spectral measure $\rho^{}_f$.
\end{coro}
Of particular relevance is the question whether the spectral measures are pure point. Here, the following holds as shown in \cite{LS}, Lemma 2.1.
\begin{lemma}\label{characterizationpp}
Let $T$ be a strongly continuous unitary representation of $\mathbb{G}$ on $\mathcal{H}$.
Then, the following assertions are equivalent for $f\in \mathcal{H}$:
\begin{itemize}
\item[(i)] The map $G\longrightarrow \mathcal{H}$, $t\mapsto T_t f$, is almost periodic in the sense that for any $\varepsilon >0$ the set $\{t \in \mathbb{G} : \|T_t f - f \|\leq \varepsilon\}$ is relatively dense in $\mathbb{G}$.
\item[(ii)] The hull $\{T_t f: t\in \mathbb{G}\}$ is relatively compact.
\item[(iii)] The map $G\longrightarrow \mathbb{C}$, $t\mapsto \langle f \mid T_t f \rangle $, is strongly almost periodic.
\item[(iv)] $\rho_f$ is a pure point measure.
\item[(v)] $f$ belongs to $\mathcal{H}_p$.
\end{itemize}
\end{lemma}
The preceding equivalence allows one to show some stability properties of $\mathcal{H}_p$. This is discussed next (see \cite{LS} as well for a proof of (a) of the following theorem).
There, we say that a \textit{measure $\mu$ is supported on a measurable set $S$} if $\mu$ of the complement of $S$ is zero.
\begin{theorem}\label{stabilitypp}
Let $T$ be a strongly continuous unitary representation of $\mathbb{G}$ on
$\mathcal{H}$.
(a) Let $C : \mathcal{H} \longrightarrow \mathcal{H}$ be continuous with $T_t C f =
C T_t f$ for each $t\in G$ and $f\in \mathcal{H}$. Then, $C$ maps $\mathcal{H}_p$ into
$\mathcal{H}_p$. If $f$ belongs to $\mathcal{H}_p$ and $\rho_f$ is supported on the subgroup $S$
of $\widehat{\mathbb{G}}$, then so is $\rho_{Cf}$.
(b) Let $M : \mathcal{H}\times \mathcal{H}\longrightarrow \mathcal{H}$ be continuous with $T_t M (f,g) = M (T_t f, T_t g)$. Then, $M $ maps $\mathcal{H}_p \times \mathcal{H}_p$ into $\mathcal{H}_p$.
\end{theorem}
\begin{proof} The statements on $C$ and $M$ are clear from the equivalence of (v) and (ii) in the Lemma~\ref{characterizationpp} and the continuity and intertwining property of $C$ and $M$. The statement on the support in (a) is proven in \cite{LS}.
\end{proof}
\medskip
We now further restrict attention to processes with second moment. Then, Theorem \ref{thm-GA} can directly be applied to give the following result (which is well known from \cite{BM,LS,MS, Gouere}).
\begin{prop} A stationary process $(N,\mathcal{H},T)$ with second moment is pure point diffractive if and only if its autocorrelation is strongly almost periodic. It has continuous diffraction if and only if the autocorrelation is null weakly almost periodic. \qed
\end{prop}
We can then combine the above results to obtain the following.
\begin{prop} For any stationary process $(N,\mathcal{H},T)$ with second moment
the mapping \;$t \mapsto \langle N(F)\mid T_t N(F)\rangle$ is
a weakly almost periodic function on $\mathbb{G}$ for all $F \in C_c(\mathbb{G})$.
The process $(N,\mathcal{H},T)$
then has pure point (resp. continuous) diffraction if and only if \;$t \mapsto \langle N(F)\mid T_t N(F)\rangle$ is
a strongly (resp. null weakly) almost periodic function on $\mathbb{G}$ for all $F \in C_c(\mathbb{G})$.
Furthermore, the diffraction is pure point if and only if for all $F\in C_c(\mathbb{G})$, $t \mapsto T_t N(F)$ is a strongly almost periodic function on $\mathbb{G}$ with respect to the norm on
$\mathcal{H}$.
\end{prop}
\begin{proof} The first statement follows from Lemma \ref{Fft-is-wap}. We now turn to the second statement. By Theorem \ref{thm-GA} and the definition of the diffraction spectrum we have pure point (resp. continuous) diffraction if and only if $\gamma$ is strongly (resp null weakly) almost periodic. Now,
from Proposition \ref{def:gamma} and a straightforward calculation we obtain for all $F,G \in C^\mathbb{R}_c(\mathbb{G})$
and all $t \in \mathbb{G}$,
\begin{equation} \label{conv2N}
G * \widetilde{F} * \gamma(t) = \gamma(G *\widetilde{T_t F})
=\langle N(G) \mid T_t N(F)\rangle\,.
\end{equation}
With $G$ set equal to $F$ we then get the required almost periodicity properties of $t \mapsto \langle N(F) \mid T_t N(F)\rangle$ from the corresponding almost periodicity properties of $\gamma$, by
our definition of these concepts. In the reverse direction, upon linearizing the
expressions $\langle N(F) \mid T_t N(F)\rangle$ we obtain \eqref{conv2N}. From the almost periodicity properties of $t\mapsto \langle N(G) \mid T_t N(F)\rangle$ we then obtain
the desired almost periodicity properties for all convolutions of the form $G^{'} \ast \widetilde{F} \ast \gamma (-t)$.
An approximate unit argument allows us to get the strong (resp. null weak)
almost periodicity of $\gamma$ from this, see \cite{SM}, Corollary 9. Here, we use that these almost periodicity properties are stable under uniform convergence.
\smallskip
The last statement is a consequence of Lemma \ref{characterizationpp}.
This finishes the proof.
\end{proof}
We finish this section with a discussion of stability properties of almost periodicity in the case of spatial processes. We will need the following proposition on continuity of composition with a fixed function. The proposition is not hard to prove and can be found in \cite{LS}.
\begin{prop}\label{continuity-of-phi} Let $\{h_n\}$ be a sequence in $L^2 (X,\mu)$ converging to $h\in L^2 (X,\mu)$. Let $\phi$ be a continuous bounded function from the complex numbers to the complex numbers. Then, $\phi \circ h_n $ converges to $\phi \circ h$.
\end{prop}
Now, we turn to the following consequence of Theorem \ref{stabilitypp} (see \cite{LS} as well).
\begin{coro} \label{coro-stabilitypp}
Let $(N,\mathcal{H},T)$ be a spatial process with $\mathcal{H} = L^2 (X,\mu)$. Let $\phi, \psi : \mathbb{C}\longrightarrow \mathbb{C}$ be bounded continuous functions and $f,g\in \mathcal{H}_p$ be given. Then, the following holds:
(a) The function $\phi \circ f$ belongs to $\mathcal{H}_p$ and the spectral measure $\rho^{}_{\phi \circ f}$ is supported in the subgroup $S$ of $\widehat{\GG}$ if $\rho^{}_f$ is supported in this subgroup.
(b) The function $(\phi \circ f) (\psi \circ g)$ belongs to $\mathcal{H}_p$. If both $\rho^{}_f$ and $\rho^{}_g$ are supported in the subgroup $S$ of $\widehat{\GG}$ then so is $\rho^{}_{(\phi \circ f)(\psi \circ g)}$.
\end{coro}
\begin{proof} By the previous proposition the maps $C: \mathcal{H}\longrightarrow \mathcal{H}$, $f\mapsto \phi \circ f$ and $M:\mathcal{H}\times \mathcal{H}\longrightarrow \mathcal{H}, M(f,g) :=(\phi \circ f) (\psi \circ g)$ are continuous. They obviously commute with the action of $\mathbb{G}$. Both (a) and the first statement of (b) follow from Theorem \ref{stabilitypp}. It remains to show the second statement of (b): Let $\mathcal{H}^{S}_p$ be the subspace of $\mathcal{H}_p$ consisting of elements whose spectral measure is supported on $S$. Then, it is not hard to see that $\mathcal{H}^{S}_p$ is a closed subspace of $\mathcal{H}$.
By (a) and the assumption on $f$ and $g$, the spectral measures of both $\phi\circ f$ and $\phi \circ g$ belong to $\mathcal{H}^{S}_p$. Moreover, $\phi\circ f$ and $\phi \circ g$ are bounded functions. Thus, it suffices to show that $\mathcal{H}^{S}_p \cap L^\infty (X,\mu)$ is an algebra. This can be shown by mimicking the proof of Lemma 1 in \cite{BL} (see \cite{LMS1} as well). For the convenience of the reader we sketch a proof:
Note first that every
eigenfunction can be approximated by bounded eigenfunctions via a simple cut-off
procedure viz if $e$ is an eigenfunction, then for an arbitrary $m>0$,
the function
\begin{equation} \label{cutoff}
e^{(m)} (x) \; := \; \begin{cases} e(x ),
& |e(x )|\le m \\ 0 \, ,
& \text{otherwise}
\end{cases}
\end{equation}
is again an eigenfunction (to the same eigenvalue). Evidently, the $e^{(m)}$ converge to $e$ in $L^2$ as
$m\rightarrow\infty$. Now, let non-zero functions $a,b\in \mathcal{H}^{S}_p\cap L^\infty (X,\mu)$ be given and choose $\varepsilon>0$ arbitrarily. As
$a$ belongs to $\mathcal{H}^S_p$, there exists
a finite linear combination $a^{'}= \sum a_i e_i$ of eigenfunctions to eigenvalues in $S$ with
\begin{equation*}
\|a - \sum a_i e_i \|_2 \; \le \; \frac{ \varepsilon}{\|b\|_\infty}\, .
\end{equation*}
By the conclusion after Eq.~(\ref{cutoff}), we can
assume that all $e_i$ are
bounded functions. Thus, in particular, $\|a^{'}\|_\infty < \infty$.
Similarly, we can choose another finite linear combination $b^{'} = \sum b_j
e_j$ of bounded eigenfunctions to eigenvalues in $S$ with
\begin{equation*}
\|b - \sum b_j e_j\|_2 \; \le \;
\frac{\varepsilon}{\|a^{'}\|_\infty} \, .
\end{equation*}
Then,
\begin{equation*}
\| a b- a^{'} b^{'}\|_2 \; \le \; \|a^{'}\|_\infty \, \| b - b^{'}\|_2 +
\|b\|_\infty \, \|a - a^{'} \|_2 \; \le \; 2\hspace{0.5pt} \varepsilon.
\end{equation*}
The product of bounded eigenfunctions to eigenvalues in $S$ is again a bounded
eigenfunction to an eigenvalue in $S$ (as $S$ is a group) and the claim follows.
\end{proof}
\section{Decomposing processes, the pure point-continuous split, and the
Gelfand construction}\label{pc:split}
In this section we first discuss how any stationary process splits
into two subprocesses, one of which has pure point spectrum and one of which has continuous spectrum. If the original process has a second moment then so have the two subprocesses and their diffraction measure will be pure point and purely continuous respectively (Theorem \ref{split}). If furthermore the original process is spatial, then its pure point part is canonically isomorphic to a spatial pure point process (Theorem \ref{prop-decomposition}). This last result is based on
the Gelfand construction of commutative Banach algebras.
\bigskip
Whenever $T$ is a strongly continuous unitary representation of $\mathbb{G}$ on the Hilbert space $\mathcal{H}$ and $U$ is a $T$ invariant closed subspace of $\mathcal{H}$ with corresponding orthogonal projection $P_U$ we can form the restriction $T_{U}$ of $T$ to $U$ and this will be a strongly continuous unitary representation as well. The orthogonal complement $U^\perp$ of $U$ in $\mathcal{H}$ is also $T$ invariant and we can form the restriction $T_{U^\perp}$ as well.
We will now focus on the special decomposition into the continuous and its point part. More precisely, we can define the continuous Hilbert space $\mathcal{H}_c$ and the point Hilbert space $\mathcal{H}_p$ by
$$ \mathcal{H}_p:=\{f\in \mathcal{H} : \rho_f \mbox{ is a pure point measure}\},\;\: \mathcal{H}_c:=\{f\in \mathcal{H} : \rho_f \mbox{ is continuous}\}. $$
Then, both $\mathcal{H}_p$ and $\mathcal{H}_c$ are closed $T$-invariant with
$$\mathcal{H}=\mathcal{H}_p \oplus \mathcal{H}_c.$$
With the projections $P_c$ and $ P_p$ onto $\mathcal{H}_p$ and $\mathcal{H}_c$ we then have $ P_c \oplus P_p = id$ and $P_* T_t = T_t P_* =: T_t^*$ for $*=p,c$. This gives the decomposition of the representation $T$ as $T = T_p \oplus T_c$ with $T_*:= P_* T$, $* = p,c$. Obviously, all spectral measures associated to $T_p$ are purely discrete and all spectral measures associated to $T_c$ are purely continuous. Accordingly, we call $T_p $ the \textit{pure point part of the representation $T$} and $T_c$ the \textit{continuous part of the representation $T$}.
\smallskip
We now turn to the case that we are given a stationary process $(N,\mathcal{H},T)$.
Of course, $T$ is then a strongly continuous unitary representation of $\mathbb{G}$ on $\mathcal{H}$. So, any $T$ invariant subspace $U$ gives rise to a decomposition of $T$. Moreover, the $\mathbb{G}$-invariance of $N$ then implies that $N$ can as well be decomposed into $N_U := P_U N$ and $N_{U^\perp} = P_{U^\perp} N$. Accordingly, the process $\mathcal{N}$ can be decomposed in to the two processes $\mathcal{N}_U := (N_U, U, T_U)$ and $\mathcal{N}_{U^\perp} := (N_{U^\perp}, U^\perp, T_{U^\perp})$ and we write $\mathcal{N} = \mathcal{N}_U \oplus \mathcal{N}_{U^\perp}$.
\smallskip
As just discussed, $T$ can be decomposed into its pure point part and its continuous part. into $N = N_p \oplus N_c$ with $N_*=P_* N$, $*=p,c$. Hence, any process $\mathcal{N}:= (N,\mathcal{H},T)$ can be decomposed into the two processes $\mathcal{N}_p:= (N_p,\mathcal{H}_p,T_p)$ and $\mathcal{N}_c:= (N_c,\mathcal{H}_c,T_c)$ which are called the \textit{pure point part} and the \textit{continuous part} of $\mathcal{N}$ respectively.
Now, assume that the original process has a second moment and associated autocorrelation measure $\gamma$ and diffraction measure $\omega = \widehat{\gamma}$. By Theorem \ref{thm-GA}, this $\gamma$ can be decomposed uniquely into its strongly almost periodic part $\gamma_{sap}$ and its null weakly almost periodic part $\gamma_{0wao}$ and the Fourier transform of $\gamma_{sap}$ is the pure point part $\omega_p$ of $\omega$ and the Fourier transform of $\gamma_{0wap}$ is the continuous part $\omega_c$ of $\omega$. It turns out that these are just the autocorrelation and diffraction measures of the pure point part and the continuous part of $\mathcal{N}$. This is the content of the next theorem.
\begin{theorem} \label{split} Let $\mathcal{N} = (N, \mathcal{H}, T)$ be a stationary process with a second moment and associated autocorrelation $\gamma$, diffraction measure $\omega$ and diffraction to dynamics map $\theta$. Let $ \gamma_{sap}$ and $\gamma_{0wap}$ be the strongly almost periodic part and the null weakly almost periodic part of $\gamma$ and $\omega_p$ and $\omega_c$ the respective Fourier transforms. Then, the point part $\mathcal{N}_p$ of $\mathcal{N}$ has a second moment with autocorrelation given by $\gamma_{sap}$ and diffraction measure given by $\omega_p$ and the continuous part $\mathcal{N}_c$ of $\mathcal{N}$ with autocorrelation given by $\gamma_{0wap}$ and diffraction measure given by $\omega_c$.
\end{theorem}
\begin{proof} By Proposition \ref{def:gamma} it suffices to show that
\begin{equation} \label{toshow} \gamma_{sap} (F\ast \tilde G) = \langle P_p N (F)\mid P_p N (G)\rangle\:\;\mbox{and}\;\: \gamma_{0wap} (F\ast \tilde G) = \langle P_c N(F)\mid P_c N(G)\rangle \end{equation}
for all $F,G\in C_c (\mathbb{G})$. By linearity it suffices to consider the case $F = G$.
Let $F_t$ be the function $T_t F = F(\cdot- t)$ and consider the function $t \mapsto \gamma (F\ast \tilde F_t)$.
We will decompose this function in a strongly almost periodic part and a null weakly almost periodic part in two ways and then conclude the desired statement from the uniqueness of the decomposition. Here are the details:
By definition of $\gamma_{sap}$ and $\gamma_{0wap}$ we can decompose this function as
$$ \gamma (F \ast \tilde F_t) = \gamma_{sap} (F \ast \tilde F_t ) + \gamma_{0wap} ( F \ast \tilde F_t)$$
with a strongly almost periodic function $t \mapsto \gamma_{sap} (F \ast \tilde F_t )$ and a null weakly almost periodic function $t\mapsto \gamma_{0wap} ( F\ast \tilde F_t)$. On the other hand by definition of $\gamma$ and the fact that $P_c$ and $P_p$ are orthogonal we also obtain the decomposition
$$ \gamma (F \ast \tilde F_t) = \langle N(F) \mid T_t N(F )\rangle = \langle P_p N(F) \mid T_t P_p N(F )\rangle + \langle P_c N(F)\mid T_t P_c N(F)\rangle.$$
Now, $t\mapsto \langle P_p N(F) \mid T_t P_p N(F)\rangle = \int (t,\xi) d \rho_{P_p N(F)}$ is strongly almost periodic as it is the inverse Fourier transform of a pure point measure and $t\mapsto \langle P_c N(F) \mid T_t P_c N(F)\rangle = \int (t,\xi) d \rho_{P_c N(F)}$ is null weakly almost periodic as it is the inverse Fourier transform of a continuous measure. From Theorem \ref{thm-GA} we conclude that
$$ \gamma_{sap} (F \ast \tilde F_t ) = \langle P_p N(F) \mid T_t P_p N(F )\rangle \;\:\mbox{and}\;\: \gamma_{0wap} ( F \ast \tilde F_t) = \langle P_c N(F)\mid T_t P_c N(F)\rangle$$
for all $t\in \mathbb{G}$. Setting $t = 0$ we obtain \eqref{toshow}.
\end{proof}
In the situation of the previous theorem we can easily see that
the diffraction to dynamics map $\theta : L^2 (\widehat{\mathbb{G}}, \omega)\longrightarrow \mathcal{H}$ can be decomposed as
\begin{equation}
\theta = \theta_p \oplus \theta_c.
\end{equation}
Here, $\theta_p$ and $\theta_c$ are the diffraction to dynamics maps of $\mathcal{N}_p$ an $\mathcal{N}_c$ respectively and and the natural decompositions $L^2 (\widehat{\mathbb{G}}, \omega) = L^2 (\widehat{\mathbb{G}}, \omega_p) \oplus L^2 (\widehat{\mathbb{G}}, \omega_c)$ and $\mathcal{H} = \mathcal{H}_p \oplus \mathcal{H}_c$ are implicitly used.
\bigskip
We now restrict attention to spatial processes. In this situation we can induce subprocesses via subalgebras of $L^\infty$. This is discussed next. We will use the main result of Gelfand theory that any commutative $C^\ast$-algebra $\mathcal{A}$ with a unit is canonically isomorphic to the algebra $C(X_\mathcal{A})$ of continuous functions on the compact space $X_\mathcal{A}$ given by the maximal ideals of $\mathcal{A}$. We will denote the space of all continuous bounded complex valued functions on the complex plane by $C_b (\mathbb{C})$.
\begin{prop} \label{prop-inducing} Let $\mathcal{N}:=(N,\mathcal{H},T)$ be a spatial process with Hilbert space $\mathcal{H}=L^2 (X,\mu)$. Let $\mathcal{A}$ be a $\mathbb{G}$-invariant subalgebra of $L^\infty (X,\mu)$ which is closed under complex conjugation, contains the constant functions and is closed with respect to the sup norm. Let $U = U (\mathcal{A})$ be the subspace of $\mathcal{H}$ generated by $\mathcal{A}$ and let $X_\mathcal{A}$ the maximal ideal space of $\mathcal{A}$.
Then, $X_\mathcal{A}$ can be equipped in a unique way with a $\mathbb{G}$ action and a $\mathbb{G}$-invariant measure $\mu_\mathcal{A}$ such that the canonical Gelfand isomorphism
$J:C(X_\mathcal{A})\longrightarrow \mathcal{A}$ extends to an unitary $\mathbb{G}$-map (also called $J$)
$$J: L^2 (X_\mathcal{A},\mu_\mathcal{A})\longrightarrow U.$$
The map $J$ is compatible with the algebraic structure in that it satisfies
\begin{itemize}
\item $J( fg) = J(f) J(g)$ for all $f\in L^\infty (X_\mathcal{A},\mu_\mathcal{A}) $ and $g\in L^2 (X_\mathcal{A}, \mu_\mathcal{A})$ and
\item $J(\phi (f)) = \phi ( J(f))$ for all $f\in L^2$ and for all continuous bounded
$\phi:\mathbb{C} \longrightarrow \mathbb{C} $.
\end{itemize}
The subspace $U$ is a $\mathbb{G}$-invariant subspace and $\mathcal{N}$ can be decomposed as
$$\mathcal{N} = \mathcal{N}_U \oplus \mathcal{N}_{U^\perp},$$
where $\mathcal{N}_U$ is isomorphic via $J$ to the spatial process $\mathcal{N}_\mathcal{A}$ with $N_\mathcal{A} = J^{-1} P_U N$, $\mathcal{H}_\mathcal{A} = L^2 (X_\mathcal{A},\mu_\mathcal{A})$. If $\mathcal{N}$ is ergodic, so is $\mathcal{N}_U$.
\end{prop}
\begin{proof} All statements are rather straightforward up to the compatibility of $J$ with the algebraic structure and the ergodicity.
\smallskip
This compatibility with algebraic structure will be discussed next: By definition, $J$ is an algebra isomorphism from $C(X_\mathcal{A})$ and $\mathcal{A}$. In particular, $J( f g) = J(f) J(g)$ for $f,g\in C (X_\mathcal{A})$. By a simple approximation argument, we then obtain $J( fg) = J(f) J(g)$ for all
$f\in L^\infty(X_\mathcal{A})$ and $g\in L^2(X_\mathcal{A})$. Here, we use that $ \{f g_n\}$ converges to $\{f g\}$ in $L^2((X_\mathcal{A})$ whenever $\{g_n\}$ converges to $g$ in $L^2(X_\mathcal{A})$ and $f$ is bounded. This shows that $J$ is compatible with multiplication.
As $J$ is an isomorphism of algebras, we also have $J (q (f)) = q( J(f))$ for any polynomial $q$ and any $f\in C (X_\mathcal{A})$. By a Stone/Weierstrass type argument, this gives $J (\phi (f)) = \phi ( J(f))$ for any $\phi \in C_b (\mathbb{C})$ and $f\in C(X_\mathcal{A})$. Consider now the set
$$\mathcal{C} :=\{f\in L^2 (X_\mathcal{A}, \mu_\mathcal{A}) : \phi (J(f)) = J(\phi (f))\;\mbox{for all $\phi \in C_b (\mathbb{C})$}\}.$$
Then, $ C(X_\mathcal{A}) \subset\mathcal{C}$ by what we have just shown. Moreover,
$\mathcal{C}$ is closed in $L^2$ by Proposition \ref{continuity-of-phi}. Thus,
$\mathcal{C}$ must agree with $L^2 (X_\mathcal{A}, \mu_\mathcal{A})$ as $C(X_\mathcal{A})$ is dense in $L^2 (X_\mathcal{A},\mu_\mathcal{A})$
(as $\{\phi (h_n)\}$ converges to $\phi (h)$ whenever $\{h_n\}$ converges to $h$).
\smallskip
It remains to show the statement on ergodicity. By definition ergodicity means that the eigenspace of $T$ to the eigenvalues $1$ is one-dimensional. This is obviously stable under the above constructions.
\end{proof}
\begin{definition} The process $\mathcal{N}_\mathcal{A}$ constructed in the previous proposition is called the subprocess of $\mathcal{N}$ induced by $\mathcal{A}$.
\end{definition}
\begin{theorem} \label{fullProcess}
Let $\mathcal{N}:=(N,\mathcal{H},T)$ be a spatial process with Hilbert space $\mathcal{H}=L^2 (X,\mu)$. Let $\mathcal{A}$ be the closure with respect to the supremum norm of the subalgebra of $L^\infty (X,\mu)$ generated by $\psi_1\circ N (F_1)\ldots \psi_n \circ N(F_n)$ with $n\in \mathbb{N}$, $F_j \in C_c (\mathbb{G})$ and $\psi_j$ continuous bounded functions from $\mathbb{C}$ to $\mathbb{C}$. Let $U = U (\mathcal{A})$ be the subspace of $\mathcal{H}$ generated by $\mathcal{A}$. Then $U$ contains the range of $N$ and the subprocess $\mathcal{N}_U$ is isomorphic to the full spatial process $\mathcal{N}_\mathcal{A}$. The process $\mathcal{N}_\mathcal{A}$ is ergodic if $\mathcal{N}$ is ergodic. Moreover,
$$\langle N_\mathcal{A} (F) \mid N_{\mathcal{A}} (G)\rangle = \langle N(F) \mid N(G)\rangle$$
holds for all $F,G\in C_c (\mathbb{G})$. In particular, $\mathcal{N}$ has a second moment if and only if $\mathcal{N}_\mathcal{A}$ has a second moment and their autocorrelations agree in this case.
\end{theorem}
\begin{proof} By definition of $\mathcal{A}$, the range of $N$ is obviously contained in the subspace $U$. Thus, $P_U N = N$ and therefore
$$ \langle N_U (F)\mid N_U (G)\rangle = \langle N (F) \mid N(G)\rangle$$
holds for all $F,G\in C_c (\mathbb{G})$. Given this all statements of the theorem follow immediately from the previous proposition.
\end{proof}
\begin{remark} The theorem says that to any (ergodic) stationary process $\mathcal{N}$ we can find an (ergodic) stationary subprocess $\mathcal{N}_U$ which is full and has the same second moment. In the situation of the theorem fullness of the original process $\mathcal{N}$ then just means that $\mathcal{N} = \mathcal{N}_{U}$.
\end{remark}
Now let $(N,\mathcal{H},T)$ be a spatial process with $\mathcal{H}=L^2 (X,\mu)$. Let
$V \subset L^2(X,\mu)$ be the closure of the linear span of products of the form
$$ \psi_1\circ N_p (F_1)\ldots \psi_n \circ N_p (F_n) ,\; n\in \mathbb{N}, \psi_j \in C_b (\mathbb{C}), F_j\in C_c (\mathbb{G}). $$
Then, $V$ is a closed invariant subspace of $L^2(X,\mu)$. Let $P_V$ be the orthogonal projection onto $V$. Then, $P_V$ commutes with $T$. Then, $(N_V, V, T_V)$ with $N_V =P_V N$ and $T_V = P_V T = T P_V$ is a stationary process. We call $(N_V,V,T_V)$ the {\it augmented point part} of the stationary process $(N,\mathcal{H},T)$. Let $V^\perp$ be the orthogonal complement of $V$ in $\mathcal{H}$. Then, $V^\perp$ gives rise to a stationary process $(N_{V^\perp},V^\perp, T_{V^\perp})$ and $N$ is the sum of these processes.
\begin{theorem} \label{prop-decomposition} Let $\mathcal{N}:=(N,\mathcal{H},T)$ be a spatial process with Hilbert space $\mathcal{H}=L^2 (X,\mu)$ and associated augmented point part $\mathcal{N}_V:=(N_V, V, T_V)$. Then the following holds:
\begin{itemize}
\item $\mathcal{N} = \mathcal{N}_V \oplus \mathcal{N}_{V^\perp}$;
\item $\mathcal{N}_V$ is isomorphic to a full spatial process;
\item $\langle P_V N(F)\mid P_V N(G) \rangle = \langle P_p N(F)\mid P_p N(G)\rangle$ and $\langle P_{V^\perp} N(F)\mid P_{V^\perp} N(G)\rangle = \langle P_c N(F)\mid P_c N(G)\rangle$ hold for all $F,G\in C_c (\mathbb{G})$.
\end{itemize}
If $\mathcal{N}$ has a second moment with autocorrelation $\gamma$ and diffraction $\omega$ furthermore the following is true:
\begin{itemize}
\item $\mathcal{N}_V$ has a second moment with autocorrelation $\gamma_{sap}$ and diffraction measure given by the pure point part of $\omega$.
\item $\mathcal{N}_{V^\perp}$ has a second moment with autocorrelation $\gamma_{0wap}$ and the diffraction measure given by the continuous part of $\omega$.
\end{itemize}
\end{theorem}
\begin{proof} Let $\mathcal{N}_p^{'} = (N_p, \mathcal{H}, T)$.
Let the algebra $\mathcal{A}$ be the closure with respect to the sup-norm of the linear span of products of the form
$$ \psi_1\circ N_p (F_1)\ldots \psi_n \circ N_p (F_n) ,\; n\in \mathbb{N}, \psi_j \in C_b (\mathbb{C}), F_j\in C_c (\mathbb{G}). $$
Then, $V$ is just the subspace generated by $\mathcal{A}$.
We can then apply the previous theorem to $\mathcal{N}_p^{'}$ and the algebra $\mathcal{A}$. This gives a decomposition $\mathcal{N}_p^{'} = \mathcal{N}_{V}^{'} \oplus \mathcal{N}_{V^\perp}^{'}$, where $\mathcal{N}_{V}^{'}$ is isomorphic to a full spatial process and
$$ \langle N_V^{'} (F), N_V^{'} (G)\rangle = \langle N_p (F), N_p(G)\rangle = \langle P_p N (F)\mid P_p N (G)\rangle $$
holds for all $F,G\in C_c (\mathbb{G})$. Here, the last equality follows by definition of $N_p$ as $P_p N$.
\smallskip
It remains to show that $P_V N (F) = P_p N(F)$ (then all remaining statements follow easily). To do so, we first note that the range of $N_p = P_p N $ is contained in the range of $U$ by construction. Moreover, by general principles the algebra $\mathcal{A}$ and hence the subspace $V$ is contained in the pure point part of $T$ \cite{BL,LMS1}. Thus, $P_V N = P_p N$ holds.
Similar remarks apply to $N_c$. Note that $P_{V^\perp} N = N - P_{V}N = N - P_p N = P_c N$.
\end{proof}
The considerations of this section and in particular the preceding theorem have the following consequence: whenever we are given a spatial process we can restrict attention to its point part and obtain a full spatial process with pure point spectrum. In particular, the whole theory developed below for full spatial processes with pure point spectrm will apply to the point part of any spatial process.
\section{Spatial processes with pure point diffraction}\label{Purepoint}
In this section we discuss spatial processes with pure point diffraction. It is these processes that we will be able to classify. We show that for these processes the two notions of pure pointedness of the spectrum defined above viz pure point diffraction spectrum and pure point dynamical spectrum actually agree. As noted in Theorem \ref{prop-decomposition}, any spatial process with second moment can be decomposed into a full spatial process with pure point diffraction and a general stationary process with continuous diffrction. Thus, the results of this section apply to the corresponding parts of any spatial process.
\bigskip
Let $(N,X,\mu,T)$ be a stationary process with pure point diffraction as discussed in Section \ref{Stochasticprocesses}. Thus, its diffraction measure $\omega$ is a pure point measure and the set of its atoms denoted by $\mathcal{S}=\mathcal{S} (\omega)$ is given by
$$\mathcal{S} =\{k\in \widehat{\mathbb{G}}: \omega (k) >0\}.$$
Let $$\theta: L^2 (\widehat{\mathbb{G}},\omega)\longrightarrow L^2 (X,m)$$
be the associated diffraction to dynamics map and $f_k := \theta (1_k)$, $k\in \mathcal{S}$.
While the subspace $\theta (L^2 (\widehat{\mathbb{G}}, \omega)$ of $L^2 (X,\mu)$ may be small in some sense, in another sense it controls the whole Hilbert space $L^2 (X,\mu)$ due to our assumption of pure point spectrum. A precise version is given next.
\begin{theorem} \label{ergodic+pure point}
Let $\mathcal{N} = (N,X,\mu,T)$ be a spatial stationary process, which is full and possesses a second moment. Then, the following assertions are equivalent:
\begin{itemize}
\item[(i)] The diffraction measure $\omega$ of $(N,X,\mu,T)$ is a pure point measure.
\item[(ii)] The representation $T$ has pure point spectrum.
\end{itemize}
In this case, the group $\mathcal{E}$ of eigenvalues of $T$ is generated by the set
$\mathcal{S}(\omega)$, and any eigenfunction is a (multiple of a) product of eigenfunctions of the form $f_k$, $k\in \mathcal{S}(\omega)$.
\end{theorem}
\begin{proof} We mimick the argument of \cite{LS} (see \cite{BL, LMS1} as well). By definition of the spectral measures $\rho_f$ given in \eqref{spectralmeasure} we have
$$ \langle N(F) \mid T_t N(F)\rangle = \int_{\widehat{\mathbb{G}}} (\gamma,t) d\rho_{N(F)} (\gamma)$$
for all $F\in C_c (\mathbb{G})$. Moreover, by definition of $\omega$ and $\gamma$ and the fact that $N$ is a $\mathbb{G}$-mapping we find
$$\langle N(F) \mid T_t N(F)\rangle = \gamma (F* \widetilde{F(T_{-t} \cdot)}) = \int_{\widehat{\mathbb{G}}} (\gamma,t) |\widehat{F}|^2 d\omega (\gamma)$$
first for real-valued $F\in C_c (\mathbb{G})$ and then by linearity for all $F\in C_c (\mathbb{G})$.
Taking inverse Fourier transforms we infer
\begin{equation}\label{ift}
\rho_{N(F)} = |\widehat{F}|^2 d\omega\;\:
\end{equation}
for all $F\in C_c (\mathbb{G})$.
\smallskip
The implication $(ii)\Longrightarrow (i)$ is now clear from \eqref{ift}. (We also saw this
earlier from the properties of $\theta$.) We next show $(i)\Longrightarrow (ii)$: As the diffraction measure is a pure point measure, the spectral measures associated to $N(F)$ are pure point measures (supported on $\mathcal{S}$) by \eqref{ift}. Thus, by part (a) of Corollary \ref{coro-stabilitypp}, the spectral measures of $\psi \circ N (F)$ for $\psi \in C_c (\mathbb{C})$, $F\in C_c (\mathbb{G})$, are then also pure point measures supported on the group generated by $\mathcal{S}$. By the fullness assumption and part (b) of Corollary \ref{coro-stabilitypp} we then infer that $T$ has pure point spectrum with eigenvalues contained in the group generated by $\mathcal{S}$.
Now, multiplying eigenfunctions $f_k$, $k\in \mathcal{S}$, we obtain eigenfunctions for arbitrary elements in the group generated by $\mathcal{S}$. This shows that the group generated by $\mathcal{S}$ is indeed the group of eigenvalues. At the same time it gives an eigenfunction for each eigenvalue. By ergodicity the multiplicity of each eigenvalue is one and these eigenfunctions are all eigenfunctions.
\end{proof}
\begin{remark} (a) There is quite some history to Theorem \ref{ergodic+pure point}: The implication $(ii)\Longrightarrow (i)$ is sometimes known as Dworkin argument. It was first established by Dworkin \cite{Dworkin} with later extensions by Hof \cite{Hof1} and Schlottmann \cite{Martin} (see \cite{Rob2,Sol2} for remarkable applications as well). In special situations equivalence was then shown by Lee/Moody/Solomyak in \cite{LMS1}. This was then generalized in various directions by Gou\'{e}r\'{e} \cite{Gouere}, Baake/Lenz \cite{BL}, Deng/Moody \cite{XR}, and Lenz/Strungaru \cite{LS} (the latter result containing all earlier ones). The above result contains all these results (provided the underlying process is real).
(b) The equivalence between diffraction and dynamical spectrum cannot hold for the complete spectrum, as discussed by van Enter/Mieskicz \cite{EM}. Indeed, \cite{EM} gives an example of mixed spectrum where the pure point part of the dynamical spectrum is completely missing in the diffraction (up to the constant eigenfunction).
\end{remark}
Due to the previous theorem we do not need to distinguish between pure point diffraction and pure point dynamical spectrum when dealing with full processes with second moments. This suggests the following definition.
\begin{definition} \label{def-pure-point-process} A spatial stationary process $ (N,X,\mu,T)$ is called \textit{pure point} if it is full, possesses a second moment and its diffraction measure is a pure point measure.
\end{definition}
If $0 \in \mathcal{S}$ in the situation of Prop.~\ref{eigenExpansionOfN(F)}
then by Lemma~\ref{fminusk} $f_0$ is a real function belonging to the
$k=0$-eigenspace of $L^2(X,m)$ and $||f_0||= \omega(0)^{1/2}$.
Since constant function ${\it 1}_X: \xi \mapsto 1$
for all $\xi \in X$ is an eigenfunction for $k=0$ and the dynamical system is ergodic,
it spans the $k=0$-eigenspace of $L^2(X,m)$. Thus $f_0 = \pm \omega(0)^{1/2}{\it 1}_X$.
Now, we have already pointed out that if $N$ is a stationary process then so is
any non-zero real multiple of it. In particular $\pm N$ are stationary processes
and one of them has the corresponding function $f_0 = \omega(0)^{1/2}{\it 1}_X$. It causes
unnecessary awkwardness later on in the discussion of phase forms to deal with the case $f_0 = -\omega(0)^{1/2}{\it 1}_X$, so wish
to normalize $N$ to avoid this situation:
\bigskip
\begin{assumption}
\label{0-eigenfunction assumption}
\begin{center}
\fbox{\parbox[c]{115mm}
{\centerline{In the pure point ergodic case
we shall always assume that}
\centerline{$f_0 = \omega(0)^{1/2}{\it 1}_X$.}}}
\end{center}
\end{assumption}
\medskip
\section{Relators and associated phase forms}\label{Cycles}
As mentioned already, our aim is to describe all stationary processes with a given pure point diffraction measure $\omega$. This description will be given in terms of a set of objects called {\it relators} arising out of the Bragg spectrum of $\omega$. While the later sections need the material of this section, it can be read independently of the previous ones.
\bigskip
Let $\widehat{\mathbb{G}}$ be a locally compact abelian group.
Let $\mathcal{S} \subset \widehat{\mathbb{G}}$ be given with $\mathcal{S} = -\mathcal{S}$ and
let $\mathcal{E}$ be the subgroup of $\widehat{\mathbb{G}}$ generated by $\mathcal{S}$.
We shall give this group the discrete topology, and in order to keep this
clear we shall usually write $\mathcal{E}_d$ instead of $\mathcal{E}$.
In the sequel we often use $m$-tuples $(k_1, \dots, k_m) \in \mathcal{S}^m$ (for various $m$) of elements of $\mathcal{S}$ or of $\mathcal{E}_d$ and we often simply denote these by bold letters ${\bf k}$. If ${\bf k} = (k_1, \dots, k_m), {\it l} = (l_1, \dots, l_n)$ then
${\bf k}{\bf l}$ is the concatenation
$$ {\bf{ k l}} = (k_1, \dots, k_m,l_1, \dots, l_n)$$
and
$$[{\bf k}] := k_1 + \dots + k_m \in \mathcal{E}_d.$$
In this way $\mathbb{S}:= \bigcup_{m=0}^\infty \mathcal{S}^m$ becomes
a monoid under concatenation, with the empty $0$-tuple $\emptyset$ as the identity element.
A {\it relator} is any $m$-tuple ${\bf k} = (k_1, \dots, k_m) \in \mathcal{S}^m$ with
$$[{\bf k}]= k_1 + \dots + k_m =0.$$
The empty tuple $\empty$ is also taken to be a relator. Thus the relators are a
subset of $\mathbb{S}$.
Let
\begin{eqnarray*}
Z_n & := & \{{\bf k} = (k_1, \dots, k_n) \in \mathcal{S}^n\,:\, [{{\bf{k}}}] =0 \} \,, n\ge 1 \,;\\
Z_0 &:= & \{\emptyset\}\,; \\
Z& := & \bigcup_{n=0}^\infty Z_n \, .
\end{eqnarray*}
Note that $Z_1= \{(0)\}$ if $0\in \mathcal{S}$, otherwise it is empty.
Concatenations of relators are relators and this makes $Z$ a submonoid of $\mathbb{S}$.
We introduce an equivalence relation on $\mathbb{S}$ by transitive extension
of the three rules:
\begin{itemize}
\item[(R1)] ${\bf k} \sim {\bf l}$ if ${\bf l}$ is a permutation of the symbols of ${\bf k}$;
\item[(R2)] if ${\bf k}= (k_1, \dots, k_m) \in \mathbb{S}$ and $0\in \mathcal{S}$ then $(k_1, \dots, k_m, 0) \sim (k_1, \dots, k_m)$ (along with the first rule, this means that when $0\in \mathcal{S}$, zeros can be dropped or added wherever they appear);
\item[(R3)] ${\bf k} \sim {\bf l}$ if ${\bf l}$ can be obtained from ${\bf k}$ by inserting removing pairs $\{k,-k\}$, $k\in \mathcal{S}$.
\end{itemize}
The second item here is related to our Assumption ~\ref{0-eigenfunction assumption},
see below.
It is easy to see that ${\bf k} \sim {\bf l}\,,\, {\bf k'} \sim {\bf l'} \Rightarrow
{\bf k}{\bf l} \sim {\bf k'}{\bf l'}$, so multiplication descends from $\mathbb{S}$ to $\mathbb{S}^\sim:= \mathbb{S}/\sim$,
whereupon $\mathbb{S}^\sim$ becomes an Abelian group under this multiplication with
$\emptyset^\sim$ as the identity element. Note that if $0 \in \mathcal{S}$ then
$0^\sim = \emptyset^\sim$. We give $\mathbb{S}^\sim$ the discrete topology, so in particular it is a locally compact Abelian group.
Since the sum $[{\bf k}] = k_1 +\dots + k_m$ of an element of $\mathcal{S}^m$ is constant on the entire
equivalence class ${\bf k}^\sim$, we see that $Z$ is the union of all equivalence classes
with component sum equal to $0$, and we obtain $\mathcal{Z} := Z/\sim$ as a subgroup of $\mathbb{S}^\sim$.
In addition, we see that there is a surjective homomorphism
$$\phi: \mathbb{S}^\sim \longrightarrow \mathcal{E}_d,\;\mbox{with}\:\; {\bf k} \mapsto [{\bf k}]$$
and its kernel is precisely $\mathcal{Z}$.
\begin{definition} The group $\mathcal{Z}= \mathcal{Z}(\mathcal{S})$ is called the {\it relator group}. The elements of its dual group
$\widehat{\mathcal{Z}}$ are called {\it phase forms}. Thus a
phase form is a group homomorphism
\[ a^*: \mathcal{Z} \longrightarrow U(1)\]
(the latter being the unit circle in $\mathbb{C}$).
\end{definition}
\begin{remark} (a) It is useful to think of relators in terms of the cycle structure of the Cayley graph of the group of eigenvalues with respect to the generators $\mathcal{S}$. In this way the group $\mathcal{Z}$ could then be seen as a a kind of abelianized homotopy group or a first cohomology group of the graph. This works in the following way:
We begin with $\mathcal{E}$ and the set $\mathcal{S}$ which satisfies $\mathcal{S} = -\mathcal{S}$
and generates $\mathcal{E}$. The Cayley graph has vertices $\mathcal{E}$ and edges consisting of all pairs
$(x, x+k)$ where $x\in \mathcal{E}$ and $k \in \mathcal{S}$. Since for each edge $(x, x+k)$ we have the
reverse edge $(x+k, x+k-k=x)$, we can think of the graph as being non-directed. Since $\mathcal{S}$
generates $\mathcal{E}$, the graph is connected. Each
${\bf k} = (k_1, \dots, k_m)$ can be thought of as an edge sequence. Given any $x\in \mathcal{E}$
we obtain a path $x,x+k_1,x+k_1+k_2, \dots, x+k_1+ \cdots + k_m$ from it. This path is a cycle if
and only if
${\bf k}$ is a relator, that is, $[{\bf k}]=0$.
There is a sort of homotopy of edge sequences, which is generated by rules corresponding
to (R1), (R2), (R3) above. The first amounts to saying that for all $x\in \mathcal{E}$ and for
all $k,l\in\mathcal{S}$, the paths $x,x+k,x+k+l$
and $x,x+l, x+l+k$ are homotopic; the second says that if $0\in \mathcal{S}$ then
the path $x, x+0$ is homotopic with the empty path from $x$; and the third
says that the path $x,x+k, x+k+(-k)$ is homotopic to the empty path from $x$. The
group $\mathcal{Z}$ may be thought of as a sort of homotopy group for cycles originating
(and terminating) at $0$.
(b) It should be noted that the homotopy group $\mathcal{Z}$ may be infinite even when the
group $\mathbb{G}$, and hence the associated graph, is finite. We shall see this in the examples
below.
\end{remark}
The phase forms play a central role in homometry problem. We note that they depend
only on $\mathcal{S}$, not on the actual values of the diffraction measure $\omega$. We shall find it convenient
to sometimes abuse the notation and write things like $a^*(k_1,\dots, k_m)$ where
we mean $a^*((k_1,\dots, k_m)^\sim)$ when using the phase form $a^*$.
Of the various ${\bf k} = (k_1, \dots, k_m )$ that can represent a given element in
$\mathcal{Z}$ there is (at least) one of minimal length $n$. This minimal
length is denoted by $\len({\bf k})$, and is called the {\it reduced length} of ${\bf k}$. Here, of course, the length of $(k_1,\ldots, k_m)$ is given by $m$.
Define
$\mathcal{Z}_n$ to be the set of elements which have reduced length less than or equal to $n$.
We have
\begin{eqnarray*}
\mathcal{Z} &=& \bigcup_{n=0}^\infty \mathcal{Z}_n,\\
\mathcal{Z}_n \mathcal{Z}_p &\subset& \mathcal{Z}_{n+p} \quad\mbox{ for all non-negative integers n and p.}
\end{eqnarray*}
We let $\mathcal{F}(\mathcal{S})$ be the Abelian group (with discrete topology) generated by symbols
$e(k)$, $k\in \mathcal{S}$, subject only to the relations\footnote{The second of these relations comes from the underlying assumption that we are dealing with {\em real} stationary processes. If this condition were dropped then these relations would also be dropped.} $e(0) =1$ (if $0\in \mathcal{S}$) and $e(-k)e(k)=1$ for all $k\in \mathcal{S}$.
\begin{prop}\label{F(S) homomorphism} The mapping
$\varphi: e(k) \mapsto (k)^\sim$ for all $k\in\mathcal{S}$ extends to an isomorphism
$\varphi: \mathcal{F}(\mathcal{S}) \longrightarrow \mathbb{S}^\sim$.
\end{prop}
\begin{proof} The existence of $\varphi$ as a surjective homomorphism is clear since $\mathbb{S}^\sim$ is the Abelian
group freely generated by the symbols $(k)^\sim$ subject only to the relations defining
$\mathcal{F}(\mathcal{S})$. These correspond precisely to $(k)^\sim (-k)^\sim = (k,-k)^\sim = \emptyset^\sim$
and $(0)^\sim = \emptyset^\sim$ when $0\in \mathcal{S}$ which hold in $\mathbb{S}^\sim$.
A typical element $e(k_1) \dots e(k_m)$ of $\mathcal{F}(\mathcal{S})$ is in the kernel of $\varphi$ if and
only if $(k_1)^\sim \dots (k_m)^\sim = (k_1, \dots, k_m)^\sim = \emptyset$. This happens only if the non-zero components
of $(k_1, \dots, k_m)$ cancel in pairs $k,-k$. Then also the corresponding terms
in $e(k_1) \dots e(k_m)$ cancel in pairs. Since also $e(0) =1$ in $\mathcal{F}(\mathcal{S})$
when $0 \in \mathcal{S}$, we see that $e(k_1) \dots e(k_m) $ reduces to the identity element.
\end{proof}
We find it convenient to identify $\mathcal{F}(\mathcal{S})$ and $\mathbb{S}^\sim$ through Prop.~\ref{F(S) homomorphism}. Having done this we obtain from $\phi:\mathbb{S}^\sim \longrightarrow \mathcal{E}_d$, defined above, the homomorphism
$$\varphi: \mathcal{F}(\mathcal{S}) \longrightarrow \mathcal{E}_d,\:\;\mbox{with}\;\: e(k) \mapsto k \,,$$
whose kernel can be identified with $\mathcal{Z}$.
Thus, we have the following
exact sequence of groups (all given the discrete topology)
\begin{equation}\label{exactSeq1}
1 \longrightarrow \mathcal{Z} \longrightarrow \mathcal{F}(\mathcal{S})
\longrightarrow \mathcal{E}_d \longrightarrow 1.
\end{equation}
Dualization then gives the exact sequence
\begin{equation}\label{exactSeq2}
1 \longleftarrow \widehat{\mathcal{Z}} \longleftarrow \widehat{\mathcal{F}(\mathcal{S}) }
\longleftarrow \mathbb{T} \longleftarrow 1,
\end{equation}
where $\mathbb{T} := \widehat\mathcal{E}_d $ is a compact Abelian group
by Pontryagin duality. Note that we get surjectivity from the continuity of
the mappings and the
compactness of these groups. In fact, surjectivity of the map $\widehat{\mathcal{F}(\mathcal{S})} \longrightarrow \widehat{\mathcal{Z}}$ is crucial for our subsequent considerations.
\begin{definition}
The mapping $a : \mathcal{S} \longrightarrow U(1)$
satisfies the $m$-{\it moment condition} if
\begin{equation} \label{mMoment}
a(k_1)a(k_2) \dots a(k_{m}) =1
\end{equation}
whenever $k_1, \dots, k_m \in \mathcal{S}$ and $k_1 + \cdots + k_m = 0$.
\end{definition}
We shall see the reason for calling these moment conditions in
\S\ref{Special}. Notice that the first moment condition is empty (and so trivially holds!) unless $0 \in \mathcal{S}$,
in which case it says that $a(0) = 1$. The second moment condition is equivalent to the statement that $a(-k) = \overline{a(k)}$ for all $k \in \mathcal{S}$. If $0\in \mathcal{S}$
then the second moment condition alone gives $a(0)a(0)=1$, so $a(0) = \pm 1$. See
Remark~\ref{0-eigenfunction assumption}. If the first moment condition is not empty and holds then $a(0)=1$ and then the $m$-moment condition includes
the $n$-moment condition for all $n<m$ since additional slots can be filled with
zeros.
Obviously, any mapping $a: \mathcal{S} \longrightarrow U(1)$ satisfying the first and second moment conditions determines a character $\mathcal{F}(\mathcal{S})$, i.e. an element
of $\widehat{\mathcal{F}(\mathcal{S})}$, via $e(k) \mapsto a(k)$ for all
$k\in \mathcal{S}$. Conversely every character of $\mathcal{F}(\mathcal{S})$ clearly determines
a mapping $a$ satisfying the first and second moment conditions. (Notice that
when $0\in\mathcal{S}$ then $e(0)$ is the identity element of $\mathcal{F}(\mathcal{S})$ and also
$a(0) = 1$.)
In the
sequel we will use these two concepts interchangeably and use
the symbol $a$ both as a mapping on $\mathcal{S}$ and as the corresponding
character on $\mathcal{F}(\mathcal{S})$.
When thought of as elements of $\widehat{\mathcal{F}(\mathcal{S})}$ they
are called {\it elementary phase forms}.
By restriction any elementary phase form $a$ determines a phase form $a^*$, that is, an element
of $\widehat{\mathcal{Z}}$. The exact sequence \eqref{exactSeq2} shows that every phase form arises
by restriction from an elementary phase form and two elementary phase forms restrict
to the same phase form if and only their ratio
is in $\mathbb{T}$, that is to say, if and only if their ratio is a character on
$\mathcal{E}_d$.
This is the essential part of determining equivalence of pure point processes.
\begin{prop} \label{characterCondition}
\begin{itemize}
\item[{(i)}] Any mapping
$a: \mathcal{S} \longrightarrow U(1)$ satisfying the first and second moment conditions
determines a unique elementary phase form and every elementary phase form
arises in this way.
\item[{(ii)}]
Any phase form is the restriction of an elementary phase form. The ratio of any two elementary phase forms giving the same phase form is an element of \,$\mathbb{T}$.
\item[{(iii)}] An elementary phase form satisfies the $m$-moment conditions
for $m= 1,\dots,n$ if and only it kills $\mathcal{Z}_n$.
\item[{(iv)}] An elementary phase form $a$ determines a character on $\mathcal{E}_d$,
i.e. lies in $\mathbb{T}$,
if and only if it satisfies all $m$-moment conditions, $m =1,2, \dots$.
\end{itemize}
\end{prop}
\begin{proof} Parts (i) and (ii) are already done above.
$\mathcal{Z}_n$ is generated by all the ${\bf k} = (k_1, \dots, k_m)$
with $k_1 + \cdots + k_m =0$, $m \le n$. An elementary phase form $a$
kills the equivalence class of $\it k$ in $\mathcal{F}(S)$ iff
\[a(k_1) \dots a(k_m) =1 \,. \]
This proves (iii). An elementary phase form $a$ kills all of $\mathcal{Z}$ iff $a^*$ is trivial
which happens iff
$a \in \mathbb{T} = \widehat{\mathcal{E}_d}$, which gives (iv).
\end{proof}
Of particular interest to us is the case where the group $\mathcal{Z}$ is generated by $\mathcal{Z}_n$ for some $n\geq 2$. This can be understood on the level of phase forms in the following way.
\begin{lemma}\label{charChar} Let $n\ge2$. Then the following assertions are equivalent:
\begin{itemize}
\item[(i)] $\langle \mathcal{Z}_n \rangle = \mathcal{Z}$.
\item[(ii)] Any mapping $a:S \longrightarrow U(1)$ satisfying the $m$-moment conditions for $1\le m \le n$ extends to a character on $\mathcal{E}_d$ (i.e. lies in $\mathbb{T}$).
\end{itemize}
\end{lemma}
\begin{proof} (ii)$\Longrightarrow$ (i): The fact that $a:S \longrightarrow U(1)$ satisfies the $m$-moment conditions for $1\le m \le n$ is equivalent to saying that the corresponding character $a$
on $\mathcal{F}(S)$ kills $\mathcal{Z}_n$, or equivalently, kills the subgroup $\langle \mathcal{Z}_n \rangle$
that it generates. But Prop.~\ref{characterCondition} says that $a$ extends to a character
on $\mathcal{E}_d$ iff $a$ kills $\mathcal{Z}$. Suppose that fact that the mapping $a:S \longrightarrow U(1)$ satisfies the $m$-moment conditions for $1\le m \le n$ is enough to sufficient to guarantee that whenever $a$ kills $\langle \mathcal{Z}_n \rangle$ it must also
kill $\mathcal{Z}$. Then $\langle \mathcal{Z}_n \rangle=\mathcal{Z}$
because $\langle \mathcal{Z}_n \rangle \subset \mathcal{Z}$, both are closed subgroups
of $\mathcal{F}(\mathcal{S})$, and the Pontryagin duality theory says that the characters can distinguish
distinct closed subgroups.
The implication (i)$\Longrightarrow$ (ii) is clear.
\end{proof}
\medskip
The preceding considerations suggest to consider
$$n_0:=n_0 (\mathcal{Z}):= \inf\{n\in \mathbb{N}: \mbox{ $\mathcal{Z}$ generated by $\mathcal{Z}_n$}\}.$$
Here, the infimum of the empty set is defined to be $\infty$.
\begin{remark} \label{2m+1 theorem} If $\mathcal{E} = \mathcal{S} + \dots + \mathcal{S}$ (with $r$ summands) then any relator can be written
as a product of relators of length at most $2r+1$ so $\mathcal{Z}_{2r+1}$ generates $\mathcal{Z}$. Thus, in this case $n_0 \leq 2 r + 1$ \cite{LM}.
\end{remark}
\smallskip
\begin{definition} The restriction of a phase form or an elementary phase form to $Z_m$ is called its $m$th-moment.
\end{definition}
\begin{prop} \label{mthMomentCharacterization}
Two elementary phase forms $a$ and $b$ give rise to the same $m$th moments if and only if their ratio $u =b/a$ satisfies the $m$th moment condition.
\end{prop}
\begin{proof} Elementary phase forms $a,b$ give rise to equal $m$th moments if and only if
$a(k_1, \dots, k_m) = b(k_1, \dots, k_m)$ whenever $k_1, \dots, k_m \in \mathcal{S}$ with
$k_1 + \cdots + k_m =0$. However, being elementary phase forms, these equations can
be written as $a(k_1) \dots a(k_m) = b(k_1) \dots b(k_m)$, which leads to the equivalent statement
that
\[u(k_1) \dots u(k_m) =1 \]
whenever $k_1, \dots, k_m \in \mathcal{S}$ with
$k_1 + \cdots + k_m =0$. This is the $m$th moment condition.
\end{proof}
Two elementary phase forms $a,b$ give rise to the same phase form $a^*$ if and only if their ratio is
an element of $\mathbb{T}$ and so satisfies all the moment conditions. By Prop.~\ref{mthMomentCharacterization} this means that all their moments are the same, so that there is no ambiguity
in speaking of the $m$th {\it moments of phase forms}.
\begin{lemma}\label{momentsCondition} Assume $\langle \mathcal{Z}_n \rangle = \mathcal{Z}$ for some $n\geq 2$. Then two phase forms agree if and only if they have the same $m$th moments for $m=1,\ldots, n$. In particular, two elementary phase forms restrict to the same phase form if and only if they have the same $m$th moments for $m=1,\dots, n$.
\end{lemma}
\begin{proof} The first statement is clear. The second statement is a direct consequence of the first statement.
\end{proof}
\begin{remark} Any of the phase forms $a$ that we are interested in from the point of view
of diffraction satisfy the first and second moment conditions by assumption.
The relevance of the previous lemma is the following. If two stationary
processes have the same pure point diffraction and their associated group of relators satisfies $\langle \mathcal{Z}_n \rangle = \mathcal{Z}$, then the processes
are isomorphic if they have the same $m$th moments for $m=1,\dots, n$ (see Cororllary~\ref{momentsCondition1} and Corollary~\ref{momentsCondition2} for precise statements).
\end{remark}
\section{From spatial stationary processes with pure point diffraction to phase factors }\label{StochasticprocessestoDiffraction}
In this section we assume that we are given an ergodic spatial stationary process with pure point diffraction.
We will discuss how this process gives rise to a diffraction measure $\omega$, a set $\mathcal{S}$ which generates a group $\mathcal{E}$, canonical eigenfunctions $f_k$, $k\in \mathcal{S}$, and
a phase form $a^*$.
In the next section we will see that $(\omega,a^*)$ in some sense uniquely determines the process.
\bigskip
Let $(N,X,\mu,T)$ be an ergodic stationary spatial process with pure point diffraction. From the considerations of Section \ref{Stochasticprocesses} we then obtain the following:
The diffraction measure $\omega$ of $(N, X,\mu,T)$ is a pure point measure and the set of its atoms, which are called Bragg peaks, is denoted by $\mathcal{S}=\mathcal{S} (\omega)$
i.e.
$$\mathcal{S} = \mathcal{S}(\omega) :=\{k\in \widehat{\mathbb{G}}: \omega (k) >0\}$$
and satisfies $\mathcal{S} = - \mathcal{S} $.
We let $\mathcal{E}$ be the subgroup of $\widehat{\mathbb{G}}$ generated by $\mathcal{S}$ and
let $\mathcal{E}_d$ denote this group when it is given the discrete topology.
We are now clearly in the situation of the previous section. In particular, there is an associated relator group $\mathcal{Z} = \mathcal{Z} (\omega)$ at our disposal. Let $$\theta: L^2 (\widehat{\mathbb{G}},\omega)\longrightarrow L^2 (X,m)$$
be the associated diffraction to dynamics map and for each $k \in \widehat{\mathbb{G}}$, let
$1_k$ be the characteristic function on the set $\{k\}$. Let $U$ be the action
of $\mathbb{G}$ acting on $L^2(\widehat{\mathbb{G}}, \omega)$ defined by
$$U_t h (k) := (k,t) h (k)$$
and observe that whenever $\omega(k) \ne 0$, $1_k$
is a $k$ eigenfunction for this action,. It is easy to see that up to scaling these are the only eigenfunctions of $L^2(\widehat{\mathbb{G}}, \omega)$.
By the orthogonality
of eigenfunctions we have,
$$\langle 1_k \mid 1_{k'} \rangle_{\widehat{\mathbb{G}}} = \omega(k) \delta_{k,k'}$$
and $$\langle F \mid G \rangle_{\widehat{\mathbb{G}}} = \sum_{k\in \mathcal{S}} F(k) G(k) \omega(k) $$
for all $F,G \in L^2(\widehat{\mathbb{G}},\omega)$.
To each element of $ \mathcal{S}$ there exists a corresponding canonical eigenfunction. More specifically, we
define $f_k$, $k\in \mathcal{S}$, by $f_k = \theta(1_k)$. Then each $f_k$ is an eigenfunction in
$L^2(X,\mu)$ for the eigenvalue $k$ and $||f_k||_2 = ||1_k||_{\widehat{\mathbb{G}}}
= \omega(k)^{1/2}$. Since $|f_k|$ is a constant function (due to ergodicity)
and $\mu$ is a probability measure, $|f_k| = \omega(k)^{1/2}$.
Then, for all $k_1, \dots, k_m \in \mathcal{S}$,
\[||f_{k_1} \dots f_{k_m}||^2 = \int_X f_{k_1} \dots f_{k_m}\overline{f_{k_1} \dots f_{k_m}}
\,{\rm d}\mu = \int_X |f_{k_1}|^2 \dots |f_{k_m}|^2 \,{\rm d}\mu = \omega(k_1) \dots \omega(k_m) \,.\]
At this point we invoke Assumption ~\ref{0-eigenfunction assumption}
which says that if $0\in\mathcal{S}$ then $f_0 = \omega(0)^{1/2} {\it 1}_X$.
\smallskip
For ${\bf k} = (k_1, \dots, k_m)$ with $k_j \in \mathcal{S}$
and $[{\bf k}]=0$ the product $f_{k_1}\ldots f_{k_m}$ is an eigenfunction to $0$ and hence (by ergodicity) a multiple of ${\it 1}_X$, which yields
\begin{equation}\label{phaseCreated}
f_{k_1} \dots f_{k_m} = a^*(k_1,\dots, k_m)\omega(k_1)^{1/2}\dots \omega(k_m)^{1/2}
\end{equation}
for some $a^*({\bf k}) = a^*(k_1,\dots, k_m) \in U(1)$.
In this way we have a mapping $a^*: Z = Z(\mathcal{S})\longrightarrow U(1)$.
It clearly respects the multiplication by concatenation on $Z$.
In view of our assumption on $f_0$ and
Lemma~\ref{fminusk} we have $f_{-k} = \overline{f_k}$, which shows this
$a^*$ is well-defined on the equivalence classes of $Z = Z(\mathcal{S})$
defined in \S\ref{Cycles}, and in this way we obtain a phase form
\begin{equation}\label{phaseFormCreated}
a^* :\mathcal{Z} \longrightarrow U(1) \,.
\end{equation}
The preceding considerations show that any pure point stationary spatial process comes with a natural measure $\omega$ and a phase form $a^*$ as summarized in the following proposition.
\begin{prop}\label{fromsptopf} Each ergodic pure point stationary spatial process $(N,X,\mu,T)$
gives rise to a pair $(\omega, a^*)$ consisting of a pure point measure $\omega$
on $\widehat{\mathbb{G}}$ characterized by
$$ \int_{\widehat{\mathbb{G}}} \widehat{F} \overline{\widehat{G}} d\omega = \langle N(F)\mid N(G)\rangle$$
for all $F,G\in C_c (\mathbb{G})$ and the phase form $a^*$ on the relator group $\mathcal{Z}=\mathcal{Z}(\omega)$ satisfying
$$ f_{k_1} \dots f_{k_m} = a^*(k_1,\dots, k_m)\omega(k_1)^{1/2}\dots \omega(k_m)^{1/2}.
$$
\end{prop}
Next we show that $\omega$ and $a^\ast$ are a complete set of invariants for the underlying process.
\section{Isomorphism of pure point processes}\label{Uniqueness}
Recall that we have introduced a notion of isomorphism between spatial processes in \S\ref{Stochasticprocesses}. In this section we show that two pure point ergodic spatial stationary processes are isomorphic (in the sense of Definition \ref{definition-isomorphism}) if they yield the same diffraction measure $\omega$ and the same phase form $a^*$.
\bigskip
We remind the reader that our concept of a pure point spatial stationary process given in Definition \ref{def-pure-point-process} entails that the process in question is full and possesses a diffraction measure (which is pure point).
\begin{theorem}\label{uniquenesstheorem} Two pure point ergodic stationary spatial processes
based on the same group $\mathbb{G}$
are isomorphic if and only if their
associated diffraction measures and phase forms are the same.
\end{theorem}
\begin{proof}
Assume that the full processes $(N,X,\mu,T)$ and $(N',X',\mu',T')$ are pure point ergodic and
have the same diffraction measure $\omega$ and the same phase form $a^*$.
Set $\mathcal{S} := \mathcal{S} (\omega) = \mathcal{S} (\omega')$. Let $f_k\in L^2 (X,\mu)$ and $f_k'\in L^2 (X',\mu')$, $k\in \mathcal{S}$, be the corresponding natural eigenfunctions. Thus, $N (F) = \sum_{k\in \mathcal{S}} \widehat{F} (k) f_k$ and $N' (F) = \sum_{k\in \mathcal{S}} \widehat{F} (k) f_k'$
for all $F\in C_c(\mathbb{G})$. Moreover, by Theorem \ref{ergodic+pure point} products of the $f_k$ and the $f_k'$ respectively provide orthonormal bases consisting of eigenfunctions of $L^2 (X,\mu)$ and $L^2 (X',\mu')$ respectively.
\smallskip
We construct an isomorphism $M$ as follows:
We define $M$ to be the unique linear map with
$$f_{k_1}\ldots f_{k_m}\mapsto f_{k_1}'\ldots f_{k_m}'.$$
This mapping is well-defined, for suppose
that $f_{k_1}\ldots f_{k_m}$ and $f_{l_1}\ldots f_{l_n}$ are linearly
dependent. Then $k_1 + \cdots + k_m = l_1 + \cdots + l_n$,
so $k_1 +\cdots + k_m - l_1 - \cdots -l_n =0$ and
\[f_{k_1}\ldots f_{k_m}f _{-l_1}\ldots f_{-l_n}= \omega(k_1)^{1/2}\ldots \omega(k_m)\omega(l_1)^{1/2}\ldots \omega(l_n)^{1/2} a^*(k_1, \dots, k_m,-l_1,\dots, -l_n) \,. \]
Using Lemma~\ref{fminusk} we obtain
\[f_{k_1}\ldots f_{k_m} = \omega(k_1)^{1/2}\ldots \omega(k_m)^{1/2}\omega(l_1)^{-1/2}\ldots \omega(l_n)^{-1/2} a^*(k_1, \dots, k_m,-l_1,\dots, -l_n)f _{l_1}\ldots f_{l_n} \,. \]
By our assumptions we obtain the same relation when the $f$s are replaced
with $f'$s, so the same linear dependence occurs in $L^2(X',\mu')$.
As products of the $f_k$ and $f_k'$ respectively give an orthogonal basis of $L^2 (X,\mu)$ and $L^2 (X',\mu')$, and as
\[||f_{k_1} \dots f_{k_m}||^2 = \omega(k_1) \dots \omega(k_m) = ||f'_{k_1} \dots f'_{k_m}||^2\,.\]
our mapping is a unitary map. In particular it is continuous.
The definition of $M$ easily shows that it intertwines $T$ and $T'$
and maps $N(F)$ to $N' (F)$ for all $F\in C_c (\mathbb{G})$. Moreover,
by definition $M$ satisfies
$$ M(f_{k_1} \ldots f_{k_n} f_{l_1}\ldots f_{l_m}) = M ( f_{k_1} \ldots f_{k_n}) M ( f_{l_1}\ldots f_{l_m})$$
for all $k_1,\ldots, k_n\in \mathcal{S}$ and $l_1,\ldots, l_m\in \mathcal{S}$. Note that both $f_{k_1} \ldots f_{k_n} $ and $M ( f_{k_1} \ldots f_{k_n}) = f_{k_1}' \ldots f_{k_n}' $ are bounded functions. Thus, we can take
linear combinations (in $L^2$) and their limits to obtain
$$ M(f_{k_1} \ldots f_{k_n} g) = M ( f_{k_1} \ldots f_{k_n}) M (g)$$
for all $g\in L^2 (X,\mu)$ and $k_1\ldots k_n\in \mathcal{S}$. In particular,
$$ M(f_{k_1} \ldots f_{k_n} g) = M ( f_{k_1} \ldots f_{k_n}) M (g)$$
holds for all $g\in L^\infty (X,\mu)$ and $k_1\ldots k_n\in \mathcal{S}$. Another approximation argument now yields $M(f g) = M(f) M(g)$ for all $f,g\in L^\infty (X,\mu)$. This proves the isomorphism.
\smallskip
Conversely suppose that $M$ is an isomorphism between the pure point ergodic stationary
spatial processes $(N,X,\mu,T)$ and $(N',X',\mu',T')$ on $\mathbb{G}$.
As $M$ is a unitary mapping, one sees from directly from the definitions that their correlation measures,
and hence their diffraction measures are the same (if second moments exist), and then that the diffraction
to dynamics mappings are related by $\theta' = M\circ \theta$. Then looking at the
eigenfunctions we have
$M(f_k) = M(\theta(1_k) = \theta'(1_k) = f'_k$. Since the $f_k$ are bounded functions,
the multiplicative property of $M$ gives
$M ( f_{k_1} \ldots f_{k_n}) = f_{k_1}' \ldots f_{k_n}' $
for all $k_1, \dots, k_m \in \mathcal{S}$. Then the definition of the phase form
\eqref{phaseCreated} and \eqref{phaseFormCreated} shows
that both processes have the same phase form.
\end{proof}
\begin{coro} \label{momentsCondition1}
Let $\omega$ be a positive symmetric pure point measure on $\widehat{\mathbb{G}}$ and $\mathcal{Z}$ the associated group of relators.
If $\langle \mathcal{Z}_n\rangle = \mathcal{Z}$, then two stationary spatial processes with diffraction $\omega$ are spatially isomorphic if and only if the first $n$ moments of their phase forms agree.
\end{coro}
\begin{proof} This is a direct consequence of the previous theorem and Lemma \ref{momentsCondition}.
\end{proof}
\begin{remark} \label{automorphism}
(a) The above theorem deals with isomorphism between two processes. We can also characterize the automorphisms of a given pure point stationary spatial point processes with Bragg peaks $\mathcal{S}$ and eigenfunctions $f_k$, $k\in \mathcal{S}$. These automorphisms are in one to one correspondence with elements from $\mathbb{T}$ in the following way: Any character $\phi \in \mathbb{T}$ yields an automorphism
of $(N,X,\mu,T)$ via $f_k \mapsto \phi(k)f_k$. Conversely any automorphism of $(N,X,\mu,T)$ must necessarily map map $f_k$ to a multiple $\phi (k) f_k$ with $\phi (k)\in U(1)$. The arising function $\phi$ must then belong to $\mathbb{T}$ as \eqref{phaseCreated} gives
$\phi(k_1) \ldots \phi(k_m) =1$
whenever $k_1 + \cdots +k_m =0$.
(b) In the previous theorem, the isomorphism $M$ between isomorphic pure point stationary
spatial processes is not unique. However, if $M_1,M_2$ are two such isomorphisms then
$M:= M_2^{-1}M_1$ is an automorphism of the first, say $(N,X,\mu,T)$.
By (a) of this remark, we conclude that the isomorphisms are unique up to elements of $\mathbb{T}$.
\end{remark}
\section{From phase forms to stationary processes with pure point spectrum: The torus approach}\label{CyclestoStochasticprocesses}
In this section we show how to construct an ergodic full stationary process with a given pure point measure as its diffraction measure $\omega$ and a given phase form $a^*$ as its associated phase form. This yields a canonical model realizing a given diffraction measure.
\bigskip
We assume that we are given a a locally compact Abelian group $\mathbb{G}$, a pure point positive symmetric backward transformable measure measure $\omega$ on $\widehat{\mathbb{G}}$. These data give rise to
$$\mathcal{S} :=\{k\in \widehat{\mathbb{G}}: \omega(k) >0\}$$
and $\mathcal{E} = \langle \mathcal{S} \rangle_{\rm group} \subset \widehat{\mathbb{G}}$. There is an associated group of relators $\mathcal{Z}$ as discussed in Section \ref{Cycles}. We are also given $a^*\in\widehat{\mathcal{Z}}$.
We are going to construct a stationary process corresponding to $(\omega, a^*)$.
From Prop.~\ref{characterCondition} we know that there is an elementary phase form
$a:\mathcal{F}(\mathcal{S}) \longrightarrow U(1)$ (i.e. a character of $\mathcal{F}(\mathcal{S})$) which restricts to $a^*$:
in other words, we can find $a$ so that
\[a^*(k_1,\dots, k_m) = a(k_1) \cdots a(k_m)\]
whenever $[{\bf k}]=0$.
(This $a$ is not unique, but the ratio of any two of such
elementary phase forms is a character on $\mathcal{E}_d$.)
Thus, we can assume without loss of generality that $a^*$ comes from an elementary phase form $a \in \widehat{\mathcal{F}(\mathcal{S})}$.
Since $\omega$ is backward transformable there is a measure $\gamma$ on
$\mathbb{G}$ so that for all $F\in C_c(\mathbb{G})$ we have
\begin{equation} \label{eighteen}
\int_{\widehat{\mathbb{G}}} |\widehat F|^2 \,{\rm d} \omega = \int_{\mathbb{G}} F*\tilde F \,{\rm d} \gamma < \infty\, .
\end{equation}
We consider $\mathcal{E}$ to have the induced topology from $\widehat{\mathbb{G}}$ and, as usual, let
$\mathcal{E}_d$ denote the same group $\mathcal{E}$ but with the discrete topology.
The dual of $\mathcal{E}_d$
is a compact Abelian group $\mathbb{T}$. We denote the Haar measure of total
volume $1$ on $\mathbb{T}$ by $l_\mathbb{T}$ and the counting measure on
$\mathcal{E}_d$ by $l_d$.
Then, \eqref{eighteen} can be reformulated as saying that
$$ (k\mapsto \widehat{F} (k) \omega (k)^{1/2}) \in \ell^2 (\mathcal{E}).$$
For $F \in C_c(\mathbb{G})$, define a
function $n_a(F)$ on $E_d$ by
\[n_a(F) = \sum_{k \in \mathcal{S}} \widehat F(k) a(k) \omega(k)^{1/2} 1_{k} \]
where we take the positive square roots and $1_k$ is the function
on $\mathcal{E}_d$ whose value at $k$ is $1$ and which takes the value $0$
everywhere else. We have
\begin{eqnarray*}
\int_{\mathcal{E}_d} n_a(F) \overline{n_a(F)} \,{\rm d} l_d &=& \sum_{k\in \mathcal{S}}
n_a(F)(k) \overline{n_a(F)(k)} = \sum_{k\in\mathcal{S}} |\widehat F(k)|^2 \omega(k)\\
&=& \int_{\widehat{\mathbb{G}}} |\widehat F|^2\,{\rm d} \omega = \int_\mathbb{G} F*\tilde F \,{\rm d} \gamma
< \infty \nonumber \, ,
\end{eqnarray*}
which implies in particular that $n_a(F) \in L^2(\mathcal{E}_d, l_d)$.
The Fourier transform provides a fundamental isomorphism between $L^2(\mathcal{E}_d, l_d)$ and
$L^2(\mathbb{T}, l_\mathbb{T})$ taking $1_{-k}$ to the character defined by $k$
on $\mathbb{T}$. We usually denote this character by $\chi_k$. Thus we obtain, by applying the Fourier transform to $n_a (F) $,
\begin{equation}\label{def-na}
N_a(F) := \sum_{k \in \mathcal{S}} \widehat F(k) a(k) \omega(k)^{1/2} \chi_k
\in L^2(\mathbb{T}, l_\mathbb{T}) \,
\end{equation}
with
\begin{equation}\label{normEquation}
\langle N_a(F) | N_a(F)\rangle =\int_\mathbb{T} N_a(F) \overline{N_a(F)} \,{\rm d} l_\mathbb{T} =
\int_{\mathcal{E}_d} n_a(F) \overline{n_a(F)} \,{\rm d} l_d = \int_\mathbb{G} F*\tilde F \,{\rm d} \gamma \, .
\end{equation}
This provides us with the mapping
\begin{equation*}
N_a \,:\, C_c(\mathbb{G}) \longrightarrow L^2(\mathbb{T}, l_\mathbb{T}) \, ,
\end{equation*}
and shows that for any compact subset $K \subset \mathbb{G}$ we have
$||N_a(F)||_2 \le (|\gamma|(K+K))^{1/2} ||F||_{\sup}$ for all
$F\in C_c^\mathbb{R}(\mathbb{G})$ with support inside $K$, which
shows that $N_a$ is continuous.
The continuous embedding of $\mathcal{E}_d \rightarrow \widehat{\mathbb{G}}$ defined by inclusion
leads to a dense homomorphism $\mathbb{G} \rightarrow \mathbb{T}$. Then each
function on $\mathbb{T}$ ``restricts" to one on $\mathbb{G}$ and in particular
we have an obvious meaning to the functions $t \mapsto \chi_k(t)$
for all $t\in \mathbb{G}$ and for all $k \in \mathcal{E}_d$. We also obtain
a natural ergodic action of $\mathbb{G}$ on $\mathbb{T}$. For $t\in \mathbb{G}$ and $\xi \in \mathbb{T}$,
\[t.\xi : k \mapsto \chi_k(t) \xi(k)\]
for all $k \in \mathcal{E}_d$.
For $F \in C_c(\mathbb{G})$ and $k\in \mathcal{E}_d$
\[\widehat F(k) \overline{\chi_k(t)} = \int_\mathbb{G} F(x) \overline{\chi_k(x)}\overline{\chi_k(t)}
\,{\rm d} l_\GG(x) = \int _\mathbb{G} (T_t F)(x) \overline{\chi_k(x)} \,{\rm d} l_\GG (x) = \widehat{T_t F}(k) \,.\]
(Here we use the invariance of the measure $l_\GG$.)
This leads by a straightforward calculation that $T_t N_a(F) = N_a(T_t F)$, which
shows that $N_a$ is a $\mathbb{G}$-equivariant map.
\smallskip
Thus given $(\omega,a^*)$ we obtain an ergodic spatial
stationary process $(N_a,\mathbb{T},l_\mathbb{T})$. The above considerations also show that $\omega$ is the diffraction measure associated to $N_a$.
We will show next that $N_a$ is full and has associated phase form $a^*$. To do so we consider the map
$$\theta : L^2 (\widehat{\mathbb{G}}, \omega)\longrightarrow L^2 (\mathbb{T},l_\mathbb{T}), \theta (h)\mapsto \sum_{k\in\mathcal{S}} h(k) a(k) \omega(k)^{1/2}\chi_k.$$
Then, $\theta$ is an isometry with $\theta (\widehat{F}) = N_a (F)$ for all
$F\in C_c(\mathbb{G})$. By uniqueness of the dynamics-to-diffraction map we infer that $\theta$ {\em is} the dynamics-to-diffraction map. This gives in particular, that
$\theta(1_k) = a(k) \omega(k)^{1/2}\chi_k$ for all $k\in \mathcal{S}$. These are the functions
$f_k$ of the previous sections. From this equality and the fact that
$a$ is a character on $\mathcal{F}(\mathcal{S})$ it follows easily that
$a^*$ is the phase form associated to $N_a$ (see \eqref{phaseCreated}
and \eqref{phaseFormCreated}). Moreover, Proposition \ref{diffractiontodynamics} now shows that
$a(k) \omega(k)^{1/2}\chi_k = \theta (1_k)$ belongs to the closure of the linear span of the $N(F)$, $F\in C_c (\mathbb{G})$, for any $k\in\mathcal{S}$. As products of these $\chi_k$ form an orthonormal basis of $L^2 (\mathbb{T},l_\mathbb{T})$ we obtain fullness of the stationary process $N_a$. We summarize the conclusions of this section.
\begin{prop} \label{inverseProblem} Let a pure point positive symmetric backward transformable measure $\omega$ on $\mathbb{G}$ be given and $a^*$ be a phase form. Then for
any choice of elementary phase form $a$ restricting to $a^*$, $(N_a,\mathbb{T},l_\mathbb{T},T)$ is an ergodic full stationary process on $\mathbb{G}$ with diffraction measure $\omega$ and phase form $a^*$.
\end{prop}
\smallskip
The pure point process constructed in Prop.~\ref{inverseProblem} depends
on the choice of the elementary phase form $a$ that we choose to represent
the phase form $a^*$.
What happens if we choose another representing elementary phase form, $b\in\widehat{\mathcal{F}(\mathcal{S})}$? We know already by Thm.~\ref{uniquenesstheorem}
that the resulting process will be isomorphic, and since both use the
same constructed dynamical system $X = \mathbb{T}$, this difference between the two
processes is in effect an automorphism. Also, Proposition \ref{characterCondition} (ii) says that the ratio $u = b / a$ is an element of $\mathbb{T}$. From the perspective of $N_a$ and $N_b$, hten $N_b$ is in effect a translation of $N_a$ ba $u\in \mathbb{T}$ i.e.
\begin{eqnarray}\label{shifting TT}
N_b(F) &=& \sum_{k\in \mathcal{S}} \widehat F(k) b(k) \omega^{1/2} \chi_k
= \sum_{k\in \mathcal{S}} \widehat F(k) a(k) \omega^{1/2} \chi_k(u) \chi_k\\
&=& \sum_{k\in \mathcal{S}} \widehat F(k) a(k) \omega^{1/2} T_{-u}(\chi_k)
= (T_{-u} N_a)(F) \,. \nonumber
\end{eqnarray}
\medskip
\begin{remark} \label{continuousCase}
If $u = b/a$ above is actually continuous with respect to the original topology on $\mathcal{E}$
then $u$, being a character on $\mathcal{E}$, lifts to a character on $\overline{\mathcal{E}}$
and then by general character theory to all of $\widehat{\mathbb{G}}$. Thus $u$ can be identified with an element of $\mathbb{G}$. Now
\eqref{shifting TT} becomes
\begin{eqnarray}\label{shifting G}
N_b(F) &=& \sum_{k\in \mathcal{S}} \widehat F(k) b(k) \omega^{1/2} \chi_k
= \sum_{k\in \mathcal{S}} \widehat F(k) a(k) \omega^{1/2} \chi_k(u) \chi_k\\
&=& \sum_{k\in \mathcal{S}} \widehat{T_{-u} F}(k) a(k) \omega^{1/2} \chi_k
= N_a(T_{-u} F) \,. \nonumber
\end{eqnarray}
This time the difference between $N_a$ and $N_b$ is a translation on $\mathbb{G}$.
This situation occurs in the periodic situation, when $\mathbb{G}$ is compact
and $\widehat{\mathbb{G}}$ is discrete.
\end{remark}
\begin{coro} \label{momentsCondition2} Let $\omega$ be a positive backward transformable pure point measure on $\widehat{\mathbb{G}} $ and $\mathcal{Z}$ the associated group of relators.
If $\langle \mathcal{Z}_n\rangle = \mathcal{Z}$, then two mappings $a,b:\mathcal{S} \longrightarrow U(1)$
define isomorphic stationary pure point processes if and only if their first through $n$th
moments are equal.
\end{coro}
\begin{proof} This is a direct consequence of
Corollary \ref{momentsCondition1}.
\end{proof}
\section{The homometry problem}\label{Homometry}
In this section we discuss a main result of the paper viz our solution to the homometry problem for pure point diffraction.
\bigskip
\begin{definition}\label{diffractionClass}
Given $\mathbb{G}$ and a measure $\omega$ on $\widehat{\mathbb{G}}$, we let $\mathcal{N}(\mathbb{G}, \omega)$ denote
the set of all spatial isomorphism classes of ergodic full stationary spatial processes with diffraction
$\omega$ satisfying Assumption \ref{0-eigenfunction assumption}.
\end{definition}
The considerations of the previous sections rather directly prove the following result, which is in effect a solution of the homometry problem
for pure point diffraction.
\begin{theorem} \label{homometrytheorem} Let $\mathbb{G}$ be a locally compact abelian group. Let $\omega$ be a positive backward transformable pure point measure on $\widehat{\mathbb{G}}$ and $\mathcal{Z} = \mathcal{Z} (\omega)$ the associated group of relators. Then, the map
$$\widehat{\mathcal{Z}}\longrightarrow \mathcal{N}(\mathbb{G}, \omega),$$
$$ a^*\mapsto [(N_a,\mathbb{T}, l_\mathbb{T}, T)],$$
where $a$ is any elementary phase form representing $a^*$,
is a bijection.
\end{theorem}
\begin{proof}
By Proposition \ref{inverseProblem}, $(N_a,\mathbb{T},l_\mathbb{T}, T)$ is an ergodic spatial stationary process with diffraction $\omega$ and phase $a^\ast$. Moreover, by Theorem \ref{uniquenesstheorem}, the map is well-defined (i.e. independent of the choice of the elementary phase form representing $a^\ast$) and one-to-one. It is surjective by Proposition \ref{fromsptopf} and Theorem \ref{uniquenesstheorem}.
\end{proof}
We note the following immediate consequence of the previous theorem and Corollary \ref{momentsCondition1} (see Corollary \ref{momentsCondition2} as well).
\begin{coro}\label{finite-coro} Assume the situation of the theorem. If $\langle \mathcal{Z}_n\rangle = \mathcal{Z}$, then two ergodic spatial stationary processes
over $\mathbb{G}$ with diffraction measure $\omega$ are spatially isomorphic if and only if their phase forms have the same $m$-th moments for $m = 1, \ldots, n$.
\end{coro}
\begin{remark} Although it is obvious from the theory of pure point spatial processes
that we have developed, it is worthwhile noting explicitly that the problem
of classification of pure point spatial processes with a given $\omega$ depends on
only the positions of the Bragg peaks, namely $\mathcal{S}$. Once one has $\mathcal{S}$, one has
the group $\mathcal{Z}$, which depends only on the relationship of $\mathcal{S}$ to the subgroup
$\mathcal{E}$ of $\widehat{\mathbb{G}}$ that it generates. And in fact this purely algebraic problem is in reality the entire homometry problem! See in \S~\ref{compactG}.2 for a $1D$ case when $\mathbb{G}$ is finite.
\end{remark}
\section{Compact Groups}\label{compactG}
Cases when $\mathbb{G}$ is compact (or even finite) are important since they arise in the study of periodic and limit periodic cases of diffractive structures, and their relative simplicity allows us to see more clearly the nature of the spatial processes that we have introduced.
In this section we first look at what happens in general when we are given that the
group $\mathbb{G}$ is compact.
We then move to some examples. These examples revolve
around the case that $\mathbb{G}$ is finite and/or around one of the most basic of all diffraction measures, namely the diffraction of $\mathbb{Z}$, $\delta_\mathbb{Z}$. We also look at an
interesting result of Gr\"unbaum and Moore that involves rational periodic diffraction
on the line. The fact that there is anything to say in this remarkably simple case
and that it seems to be very difficult to extend their results to a two dimensional setting is a bit of
a testimony to the difficulties that are inherent in the homometry problem.
\subsection{The compact setting}\label{The compact setting}
We now look at the method of \S\ref{CyclestoStochasticprocesses} in this compact case.
We assume that $\omega$ is positive, pure point,
centrally symmetric ($\omega(-k) = \omega(k)$, for all $k\in \mathcal{S}$), and backward transformable. Let $\mathcal{S}$ be the support of $\omega$. It is convenient to assume
that $\mathcal{S}$ generates $\widehat{\mathbb{G}}$ as a group, for otherwise we may use $\langle \mathcal{S}\rangle$ and $\widehat{\langle \mathcal{S}\rangle}$ instead. Then $\mathcal{E} = \widehat{\mathbb{G}} = {\mathcal{E}}_d$
and the carrying space for our spatial process is
$$X= \mathbb{T} := \widehat{\mathcal{E}_d} = \mathbb{G}.$$
When we wish to distinguish two roles of $\mathbb{G}$, first as a group of translations and second as the set of states of dynamical system, we shall use notation
like $t$ and $\xi = \xi_t$ respectively. Moreover, for any character $k$ on $\mathbb{T}$ (i.e. $k\in \widehat{\mathbb{G}}$) we will denote the map
$$L^2 (X)\longrightarrow \mathbb{C}, \; F\mapsto \widehat{F} (k)$$
by $k$ as well.
When we come to the transformability into the autocorrelation measure $\gamma$, we require that for all $F\in C_c(\mathbb{G}) = C(\mathbb{G})$,
\begin{equation}
\label{transformability-assumption} \gamma(F*\tilde F) = \int_{\widehat{\mathbb{G}}} |\hat F|^2 d\omega =
\sum_{k \in \widehat{\mathbb{G}}} \omega(k) |\hat F|^2(k)
= \sum_{k \in \widehat{\mathbb{G}}} \omega(k) \widehat{F*\tilde F}(k).
\end{equation}
Thus the autocorrelation Fourier dual to $\omega$ is given by
\begin{equation} \label{definition-gamma}
\gamma := \sum_{k\in \mathcal{S}} \omega(k)\;k = \sum_{k\in \widehat{\mathbb{G}}} \omega(k)\;k \,
\end{equation}
provided that this is indeed a measure. Since $\mathbb{G}$ is compact, and $\gamma$
is supposed to be a Borel measure, this measure must be finite.
Assuming transformability we can, for any character $a:\mathcal{F}(\mathcal{S}) \longrightarrow U(1)$, form the associated
stationary process
\[N_a:C_c(\mathbb{G}) \longrightarrow L^2(X,\mu) \, ,\]
namely,
\[N_a(F) = \sum_{k\in\widehat{\mathbb{G}}} \hat F(k) a(k) \omega^{1/2}(k) \, k \]
for all $F\in C(\mathbb{G})$. Note that $N_a (F)$ belongs indeed to $L^2 (X,\mu)$ due to
$\eqref{transformability-assumption}$.
\smallskip
Our interpretation is that $N_a$ represents some sort of density on the space
$\mathbb{G}$. Each point $\xi_t \in X$ represents an instance of the yet unspecified
structure which can be paired with elements of $C_c (\mathbb{G})$ to give
\begin{equation}
\label{measureInterpretation}
\langle \xi_t, F\rangle = N_a(F)(\xi_t) = N_a (F) (t),
\end{equation}
where the left hand side needs to be given an interpretation. It turns out that such an interpretation can be given easily if we make the finiteness assumption that
\begin{equation}\label{finiteness-assumption}
\omega \in L^1(\widehat{\mathbb{G}},l_{\widehat{\mathbb{G}}}).
\end{equation}
Note that this assumption immediately implies that $\omega$ is backward transformable as it yields that $\gamma$ given in $\eqref{definition-gamma}$ is indeed a measure.
Given $\eqref{finiteness-assumption}$, we can
define the function $\rho_a \in L^2 (X)$ by
\begin{equation}\label{densityDefined}
\rho_a:= \sum_{k\in \widehat{\mathbb{G}}} a(k) \omega(k)^{1/2} \overline k
\end{equation}
by standard theory of Fourier series. Then, a short calculation invoking unitarity of the Fourier transform (see e.g. \eqref{def-na} as well) gives
\begin{eqnarray}\label{processAsMeasure}
N_a(F)(t) &=&\sum_{k\in \widehat{\mathbb{G}}} \hat F(k) a(k) \omega(k)^{1/2} (k,t )
\nonumber \\
&=& \int_\mathbb{G} T_t\rho_a(x) F(x) dl_\GG(x).
\end{eqnarray}
Note that we have indeed pointwise (in t) existence of $N_a (F) (t)$ as both $\widehat F$ and $\omega$ are square summable i.e. belong to $\ell^2 (\mathcal{E}_d)$. Comparing with $\eqref{measureInterpretation}$, we find that we can identify
$\xi_t$ with the $L^2$ function $T_t \rho_a$ and the pairing between $ \xi_t$ and $C_c (\mathbb{G})$ with ordinary integration.
When can this `density'
be interpreted as a measure $\rho_a$ on $X = \mathbb{G}$?
The idea is that for some $\xi_0 \in X$ (which we can take to be $0 \in \mathbb{G}$) the structure is $\rho_a$ and that as we translate around the translate $T_t\rho_a$ represents the structure at $\xi_t$. Usually one would have to assume by ergodicity the denseness of this orbit, but in the compact case as we have it here, the orbit is the entire space $X$. This entails that
\begin{equation}
\langle T_t \rho_a, F\rangle = N_a(F)(t).
\end{equation}
where
and then
$T_t\rho_a = \sum_{k\in \widehat{\mathbb{G}}} (k,t) a(k) \omega(k)^{1/2} \overline k$.
This gives \eqref{measureInterpretation} when we treat $\rho_a$ as a measure.
As $\mathbb{G}$ is compact the $L^2$-function $T_t \rho_a$ belongs to $L^1$ as well and we can consider it to be the measure
$$ T_t \rho_a dl_\GG.$$
Thus, in this situation we can realize the process as a measure process.
If $\sum \omega(k)^{1/2} <\infty$, then $\rho_a$ will even be a continuous function.
If $\sum \omega(k)^{1/2} =\infty$, the situation becomes different. Such a situation can easily arise:
it suffices to find $a$ and $\omega$ such that $\rho_a$ does not belong to $L^\infty (X)$.
Let for example $\mathbb{G}$ be compact admitting a function $\rho\in L^2 (\mathbb{G})\setminus L^\infty (\mathbb{G})$. As $\rho$ belongs to $L^2$, we
can expand $\rho$ in a Fourier series
$$\rho = \sum_ k a(k) \omega(k)^{1/2} \overline{k}$$
with $a(k)\in U(1) $ and $\omega (k)\geq 0$ with $ \sum \omega (k) < \infty$.
The diffraction of $\rho$ is seen directly to be equal to $\omega$.
For more on interpreting $N_a$ as some sort of measure induced density
on $\mathbb{G}$ see \S\ref{ssp-measures}.
\section{Examples around the diffraction measure $\omega =\delta_\mathbb{Z}$}
In general the number of solutions to the inverse problem is vast, even for
the most basic diffraction patterns. In this section we consider the most famous
example of a diffraction pattern, the diffraction of the integers. More properly
this is the diffraction of the Dirac comb $\delta_\mathbb{Z}= \sum_{x\in\mathbb{Z}} \delta_x$
which represents a point density of one at each of the integers. Its diffraction is also
$\omega =\delta_\mathbb{Z}$.
Crystallographers have long relied on additional information (like the periodicity of crystals
and their known constituents)
to find solutions to the inverse problem.
We shall see that by imposing periodicity and/or imposing arithmetic conditions on the
solutions, the inverse problem we can do the same. One of the advantages
of our approach, that applies to all locally compact Abelian groups, is that it also applies
to compact and to finite groups. Although our object of attention is
$\delta_\mathbb{Z}$, much of this section is devoted to cases in which $G$ is finite
and to showing what the theory then looks like and tells us. We also deal with the case $G=U(1)$.
\medskip
In the case of diffraction
from a periodic set or periodic crystal, it is customary to compute the diffraction directly from
the originating measure describing the density $\rho$, by taking the square-absolute value
of the Fourier transform of $\rho$. In the case of $\delta_\mathbb{Z}$ this approach gives the answer
directly as a consequence of the Poisson summation formula. This also works directly for
finite groups.
We start with the pair of dual groups $\frac{1}{M}\mathbb{Z}/\mathbb{Z}$ and $\mathbb{Z}/M\mathbb{Z}$
where $M$ is a positive integer. The duality is most conveniently written down in the form
\[ (k,x) = e^{2 \pi i kx} \]
for $x \in \frac{1}{M}\mathbb{Z}/\mathbb{Z}$ and $k \in \mathbb{Z}/M\mathbb{Z}$. In the context of this paper, one of these
groups plays the role of $\mathbb{G}$ and the other of $\widehat{\mathbb{G}}$. Most of the time
it will be $\mathbb{G} = \frac{1}{M}\mathbb{Z}/\mathbb{Z}$ and $\widehat{\mathbb{G}}=\mathbb{Z}/M\mathbb{Z}$. To help clarify things, elements of $\widehat{\mathbb{G}}$ will be written in the form $\chi_x$ or $\chi_k$ accordingly.
The idea here is to use the fact that the diffraction $\delta_\mathbb{Z}$ is periodic and then
represent it in $\mathbb{Z}/M\mathbb{Z}$ and look for all solutions for an originating distribution $\rho$
in $\frac{1}{M}\mathbb{Z}/\mathbb{Z}$. The resulting distribution will then be interpreted as a periodic
(modulo $\mathbb{Z}$) measure
on $\frac{1}{M}\mathbb{Z}$. In this way we pick up periodic solutions to the inverse problem. Alternatively
we could do things the other way around and consider $ \delta_\mathbb{Z}$ represented in
$\frac{1}{M}\mathbb{Z}/\mathbb{Z}$
and find solutions to the inverse problem in $\mathbb{Z}/M\mathbb{Z}$ which we could interpret as periodic
solutions in $\mathbb{Z}$. However, we shall see that does not lead to new solutions
to the inverse problem.
We note that for finite $\mathbb{G}$ and $\widehat{\mathbb{G}}$ the Haar measure we want to use
is counting measure
normalized to total measure $1$ by dividing by the order of the group. For
$\chi_k \in \mathbb{Z}/M\mathbb{Z}$, simple calculation gives the expected result that
$\widehat{\chi_k} = \delta_k$ (note that $\chi_k$ is a function on $\mathbb{G}$ and
its Fourier transform is a function on $\widehat{\mathbb{G}}$).
\smallskip
Since the groups are finite, there is no complication in computing diffraction.
For a density or distribution of density $\rho$ on $\mathbb{G}$, the autocorrelation
is $\gamma = \rho*\widetilde \rho$ and the diffraction is $\omega = (\rho*\widetilde{\rho})\widehat{\phantom{x}}= | \widehat \rho |^2$.
\begin{remark} We should note that in general the approach via the square-absolute value of the Fourier
transform is not applicable to aperiodic structures.
As A. Hof has shown \cite{Hof1} and \cite{Hof2}, in the study of
aperiodic crystals one usually meets distributions of density that are not Fourier transformable.
\end{remark}
\medskip
Our approach is to follow the lines laid out above and for given diffraction $\omega$
and elementary phase form $a$ define the associated density via
\eqref{densityDefined}:
\begin{equation}
\rho_a:= \sum_{k\in \widehat{\mathbb{G}}} a(k) \omega(k)^{1/2} \overline k \,,
\end{equation}
This is a finite sum and there are no questions about convergence or ambiguities
about how to interpret it.
\subsection{Specific values of $M$}
\subsubsection{The case $M=1$}
This seems too trivial to consider, but it is interesting nonetheless.
$\mathbb{G}= \frac{1}{1}\mathbb{Z}/\mathbb{Z}$, $\widehat{\mathbb{G}} =\mathbb{Z}/\mathbb{Z}$,
$\omega = \delta_0$. Here $\mathcal{S} = \{0\} \in \mathbb{Z}/\mathbb{Z} $ and
$\mathcal{F}$ is generated by $e(0)=1$. So
$\mathcal{F}$ is the trivial group and its only character is the trivial character. Also $\mathcal{Z} = \mathcal{F}$.
The density measure $\rho$ defined by the trivial character $a$ is
$a(0)\omega(0)^{1/2}\overline{\chi_0} = \chi_0$ which is simply the identity function on $\mathbb{G}$.
Thus we obtain the expected solution to the inverse problem: $\widehat{\chi_0} = \delta_0$
whose absolute square is again $\omega = \delta_0$. The function $\chi_0$ is the identity function
on $\frac{1}{1}\mathbb{Z}/\mathbb{Z}$, which is also the measure $\delta_0$ on $\mathbb{G}$.
\subsubsection{The case $M=2$}
(1) \quad $\mathbb{G} =\frac{1}{2}\mathbb{Z}/\mathbb{Z}$, $\widehat{\mathbb{G}}=\mathbb{Z}/2\mathbb{Z}$,
$\omega = \delta_0 + \delta_1$. The two elements of
$\widehat{\mathbb{G}}$ are $\chi_0$ and $\chi_{1}$. We have $\mathcal{S} = \{0, 1\} =\widehat{\mathbb{G}}$ and $\mathcal{F}$
is generated by $e(0) =1 $ and $e(1)$, with $e(1)^2=1$ since $-1 \equiv 1 \mod 2\mathbb{Z}$.
Thus $\mathcal{F}$ is cyclic of order $2$. There are only two characters on $\mathcal{F}$, i.e. elementary
phase forms,
$a_+:e(1) \mapsto 1$ and $a_-:e(1) \mapsto -1$.
The exact sequence
\[1 \longrightarrow \mathcal{Z} \longrightarrow \mathcal{F} \longrightarrow \mathcal{E} = \widehat{\mathbb{G}} \longrightarrow 1\, ,\]
comes from the mapping $e(k) \mapsto k$ which here reads
$e(0) \mapsto 0$ and $e(1) \mapsto 1$. Thus the kernel $\mathcal{Z}$ is trivial. The significance
of this is that there is only one phase form (trivial!) and thus the two solutions we derive
from the two elementary phase forms will be the same
up to translation. In detail, we have, using $a_+$
\[\rho_+ = a_+(0) \omega(0)^{1/2}\overline{\chi_0} + a_+(1)\omega(1)^{1/2}\overline{\chi_{1}}
= \chi_0 + \chi_{1}\,.\]
Note that this is a function on $\mathbb{G} = \frac{1}{2}\mathbb{Z}/\mathbb{Z}$. Its values
are $\rho_+(0) =2, \rho_+(1/2) = 0$.
This is a nice transformable measure and its Fourier transform is the measure
is $\delta_0 + \delta_1$ on $\widehat{\mathbb{G}} = \mathbb{Z}/2\mathbb{Z}$.
Its absolute square is $\omega$, as we expected.
Similarly, using $a_-$ we obtain
\[\rho_- = a_-(0) \omega(0)^{1/2}\overline{\chi_0} + a_-(1)\omega(1)^{1/2}\overline{\chi_{1}}
= \chi_0 - \chi_{1} \,.\]
Its values are $\rho_-(0) =0, \rho_-(1/2) =2$,
so the density distribution $\rho_-$ on $\mathbb{G}$ is just translation by $1/2$ of $\rho_+$
found in the case of $a_+$.
The Fourier transform is $\delta_0 - \delta_1$
and again the absolute square of this is $\omega$.
\bigskip
(2) \quad Next we reverse the roles of $\mathbb{G}$ and $\widehat{\mathbb{G}}$: $\mathbb{G} =\mathbb{Z}/2\mathbb{Z}$, $\widehat{\mathbb{G}} = \frac{1}{2}\mathbb{Z}/\mathbb{Z}$,
$\omega = \delta_0$. Notice that $\widehat{\mathbb{G}}$ represents the reduction
of the integers and the half-integers modulo $\mathbb{Z}$. The reduction of
the diffraction $\delta_\mathbb{Z}$ modulo $\mathbb{Z}$ is thus supported only on the
$\mathbb{Z}$ part which is represented by the class of $0$ modulo $\mathbb{Z}$.
Following the general discussion of compact groups we should,
at this point, replace $\widehat{\mathbb{G}}$ by the subgroup $\langle \mathcal{S} \rangle = \{0\}$.
This then reduces us to the case $M=1$ already discussed.
\bigskip
(3) \quad It is interesting to consider
$\mathbb{G} =\mathbb{Z}/2\mathbb{Z}$ and $\widehat{\mathbb{G}} = \frac{1}{2}\mathbb{Z}/\mathbb{Z}$ with the diffraction
$\omega = \delta_0 + \delta_{1/2}$. This may be considered as the reduction
modulo $\mathbb{Z}$ of $\delta_\mathbb{Z} + \delta_{\frac{1}{2} + \mathbb{Z}}$. The two elements of
$\widehat{\mathbb{G}}$ are $\chi_0$ and $\chi_{1/2}$. The discussion parallels that in (1). We have $\mathcal{S} = \{0, 1/2\}$ and $\mathcal{F}$
is generated by $e(0) =1 $ and $e(1/2)$ with $e(1/2)^2=1$ since $-1/2 \equiv 1/2 \mod \mathbb{Z}$.
Thus $\mathcal{F}$ is cyclic of order $2$. There are only two characters on $\mathcal{F}$, i.e. elementary
phase forms,
$a_+0:e(1/2) \mapsto 1$ and $a_-:e(1/2) \mapsto -1$.
The exact sequence
\[1 \longrightarrow \mathcal{Z} \longrightarrow \mathcal{F} \longrightarrow \mathcal{E} = \widehat{\mathbb{G}} \longrightarrow 1\, ,\]
comes from the mappinng $e(k) \mapsto k$ which here reads
$e(0) \mapsto 0$ and $e(1/2) \mapsto 1/2$. Thus again the kernel $\mathcal{Z}$ is trivial
and the two solutions we derive will
be the same
up to translation. Using $a_+$
\[\rho_+ = a_+(0) \omega(0)^{1/2}\overline{\chi_0} + a_+(1/2)\omega(1/2)^{1/2}\overline{\chi_{1/2}}
= \chi_0 + \chi_{1/2} \,.\]
Thus the distribution on $\mathbb{Z}/2\mathbb{Z}$ is $\rho_+(0) =2, \rho_+(1)=0$.
Its Fourier transform is the measure $\delta_0 + \delta_{1/2}$, whose absolute square
is $\omega$.
Going back to $\mathbb{Z}$ we have $2\delta_{2\mathbb{Z}}$ on $\mathbb{Z}$ whose diffraction is
$\frac{1}{2}(2 \delta_{\frac{1}{2}\mathbb{Z}}) = \delta_\mathbb{Z} + \delta_{\frac{1}{2} +\mathbb{Z}}$, which
agrees with what we obtained after removing all periodicity.
Similarly, using $a_-$ we obtain
\[\rho_- = a_-(0) \omega(0)^{1/2}\overline{\chi_0} + a_-(1/2)\omega(1/2)^{1/2}\overline{\chi_{1/2}}
= \chi_0 - \chi_{1/2} \,,\]
and the distribution on $\mathbb{Z}/2\mathbb{Z}$ is
$\rho_-(0) = 0, \rho_-(1) = 2$.
Then $\widehat{\rho_-} = \delta_0 - \delta_{1/2}$. Taking absolute squares we again
obtain the diffraction $\omega$.
Again going back to $\mathbb{Z}$, this time we have $2\delta_{2\mathbb{Z} +1}$ on $\mathbb{Z}$ whose diffraction is
$\delta_\mathbb{Z} - \delta_{\frac{1}{2} +\mathbb{Z}}$, which
agrees with what we obtained after removing all periodicity.
\bigskip
\subsubsection{The case $M=3$}
Here something more interesting happens. $\mathbb{G} = \frac{1}{3}\mathbb{Z}/\mathbb{Z}$, $\widehat{\mathbb{G}} = \mathbb{Z}/3\mathbb{Z}, \omega=\delta_0 + \delta_1+\delta_2$.
The three elements of
$\widehat{\mathbb{G}}$ are $\chi_0,\chi_{1},\chi_{2}$, with $\chi_k(j/3) = (k,j/3) = \zeta_3^{kj}$,
where $\zeta_3 =e^{2 \pi i /3}$. We have $\mathcal{S} = \{0, 1,2\} =\widehat{\mathbb{G}}$ and $\mathcal{F}$
is generated by $e(0) =1,e(1),e(2) $, with $e(2)= e(1)^{-1}$
since $-1 \equiv 2 \mod 3\mathbb{Z}$.
Thus $\mathcal{F} \simeq \mathbb{Z}$ with $e(1)$ as a (multiplicative) generator.
There are infinitely many characters on $\mathcal{F}$, i.e. elementary
phase forms, one for each $u \in U(1)$:
$a_u: e(1)^n \mapsto u^n \in U(1)$ and $e(2)^n\mapsto u^{-n}$ for all $n \in \mathbb{Z}$.
The exact sequence
\[1 \longrightarrow \mathcal{Z} \longrightarrow \mathcal{F} \longrightarrow \mathcal{E} = \widehat{\mathbb{G}} \longrightarrow 1\, ,\]
comes from the homomorphism that maps $e(1)^k \mapsto k \mod 3$ (which includes $
e(1)^{-k} = e(-1)^k\mapsto -k \mod 3$).
The kernel $\mathcal{Z} \simeq 3\mathbb{Z}$ via the isomorphism $\mathcal{F} \simeq \mathbb{Z}$.
For each $u\in U(1)$ we obtain
\begin{eqnarray}
\rho_u &=& a_u(0) \omega(0)^{1/2}\overline{\chi_0} + a_u(1)\omega(1)^{1/2}\overline{\chi_{1}} +
a_u(2)\omega(2)^{1/2}\overline{\chi_2}\\
&=& 1 \chi_0 + u\chi_{-1} + \overline u \chi_1\,.\nonumber
\end{eqnarray}
Thus the distribution on $\frac{1}{3}\mathbb{Z}/\mathbb{Z}$ is
$\rho_u(0) = 1+u+\overline u$,
$\rho_u(1/3) = 1+ u\zeta_3^2 + \overline u \zeta_3$, and
$\rho_u(2/3) =1 + u\zeta_3 + \overline u \zeta_3^2\,.$
Since $\widehat{\chi_k} = \delta_k$
we obtain $\widehat{\rho_u} = \delta_0 + u \delta_2 + \overline u \delta_1$.
Its absolute square is $\omega$ as required.
We note that replacing $u$ by $\zeta_3u$ or $\zeta_3^2u$ changes
the distribution $\rho$ by translation, which corresponds to the fact
that the elementary phase forms map $3:1$ to phase forms.
Thus there is a one parameter family of solutions to the inverse problem
for $\omega$, which are supported on $\mathbb{Z}$ and periodic of period $3$.
\smallskip
\subsubsection{Other cases of $M$}
If we look at the inverse problem to $\delta_0 + \cdots + \delta_{M-1}$
in $\frac{1}{M}\mathbb{Z}/\mathbb{Z}$ the situation evolves in a natural way.
At $M=4$ the value $2$ leads to
$e(2)^2=1$, and $e(1) = e(3)^{-1} \mapsto u \in U(1)$ is arbitrary, so we end up with solutions
$\chi_0 + u\chi_{1/4} \pm \chi_{1/2} + \overline u\chi_{3/4}$.
At $M=5$ we are get two free parameters $u,v \in U(1)$, one relating to the pair
$1,4$ and one to the pair $2,3$ (integers taken modulo $5$), and the solutions are
$\chi_0 + u\chi_{1/5} +v \chi_{2/5} + \overline v\chi_{3/5} +\overline u \chi_{4/5}$.
The results for higher $M$ follow a similar pattern.
\subsection{What happens when $\mathbb{G} = U(1), \widehat{\mathbb{G}} = \mathbb{Z}$?}\label{G=U(1)}
We again consider the diffraction $\delta_\mathbb{Z}$, but now use $\mathbb{Z}$ as $\widehat{\mathbb{G}}$
and $U(1)$ as $\mathbb{G}$.
Then $\mathcal{S} = \mathbb{Z}$ and $\mathcal{F}$ is generated by $\{e(k) :k \in \mathbb{Z} \}$ with the single
set of relations $e(-k) = \overline{e(k)}$ (along with $e(0)=1$). Then we have elementary
phase forms $a:\mathcal{F} \longrightarrow U(1)$ with $a(k) \in U(1)$ arbitrary except that
$a(0) =1, a(-k) = \overline{a(k)}$.
Each such $a$ leads to a formal solution
\[ \rho_a=\sum_{k\in\mathbb{Z}} a(k) \omega(k)^{1/2} \overline{\chi_k}
= \sum_{k\in\mathbb{Z}} a(k) e^{-2 \pi i k.(\cdot)} \, ,\]
This is simply a formal Fourier series on $\mathbb{Z}$. However, given that for all $k$,
$|a(k)|=1$, there is no convergence as a Fourier series. This is expected
since in the simplest case when $a(k)=1$ for all $k$ its Fourier transform is the
unbounded measure $\delta_\mathbb{Z}$.
However, the solution for $N_a$:
\[N_a(F) = \sum_{k\in\mathbb{Z}} \hat F(k) a(k) \omega(k)^{1/2} \, k
= \sum_{k\in\mathbb{Z}} \hat F(k) a(k) \, k \]
for all $F\in C(U(1))$ is well-defined as an $L^2$-function, and we can learn something from it.
In the case that $a$ is the trivial elementary phase form we obtain
\[N_a(F)(t) = \sum_{k\in\mathbb{Z}} \hat F(k) (k,t) = F(-t) = \delta_0(T_t F)\,.\]
Thus the associated density is $\delta_0$, which is what we expect. Recall
\S\ref{The compact setting} here.
The compact group $\mathbb{T}$, which is $U(1)$ here, is viewed both as a dynamical system
with states $\xi_t$
and as a group of translations $t\in U(1)$ which translate the states about. $N_a(F)(t)$ is something that gives the effect of the process on the function $F$ at the state $\xi_t$.
Now let $K=-K$ be a finite symmetric subset of $\mathbb{Z}\backslash \{0\}$ and let
$a$ be an elementary phase form which satisfies $a(k) =1$ for all $k \notin K$
and $a(k) = \overline{a(-k)} \in U(1)$ arbitrary for all $k\in K\cap \mathbb{Z}_+$. Now for
$F\in C(U(1))$ we have
\[N_a(F)(t) = \sum_{k\in\mathbb{Z}} \hat F(k) (k,t) + \sum_{k\in K} (a(k) -1) \hat F(k) (k,t) \,,\]
for all $t\in U(1)$. Following the same reasoning as in \eqref{processAsMeasure}
we arrive at $N_a(F)(t) = \delta_0(T_t F) + \rho(T_tF)$ where
\[\rho = \sum_{k\in K} (a(k)-1) \overline{\chi_k}\,.\]
Remarkably the density $\delta_0 + \sum_{k\in K} (a(k)-1) \overline{\chi_k}$ on $U(1)$
is indeed a solution to the inverse
problem for $\delta_0$, as can be seen by directly computing its autocorrelation
and taking its Fourier transform. This shows the ubiquity of quite reasonable looking solutions
to the problem.
\subsection{A simple periodic case with imposed arithmetic conditions} \label{simplePeriodic}
In this section we
consider the very simple situation of
the diffraction from the integers when they are weighted periodically with
period $M$ by a set $w_0, \dots, w_{M-1}$ of real numbers. In this case the
diffraction is also periodic, and we will assume that the ambient space
is $\mathbb{G} = \mathbb{Z}/M\mathbb{Z}$ and the diffraction occurs in its dual, $\widehat{\mathbb{G}} \simeq \frac{1}{M}\mathbb{Z}/\mathbb{Z}$,
say
\begin{equation}\label{finiteDiffraction}
\omega = \sum_{k\in \mathbb{Z}/M\mathbb{Z}} \omega(k) \delta_k \,.
\end{equation}
As we noted above, making the assumption that $\mathbb{G}= \mathbb{Z}/M\mathbb{Z}$ is not entirely innocent
since it puts considerable limitations on the kinds of solutions available -- namely
that they are periodic of period $M$ and the distribution of intensity is entirely confined to integers,
neither of which is mandatory. Taking a larger ambient space will in general produce more
solutions (many more!).
Equation \eqref{densityDefined} defines the density on $\mathbb{G}$ arising from
$\omega$ and the elementary phase form $a$. Since its diffraction is to be
$\omega$, this leads to the $M$ equations
\begin{equation}\label{weightEquations}
\rho_a(x) =\sum_{k\in \mathcal{S}} a(k) \omega(k)^{1/2} \overline{ (k,x)} = w_x, \qquad x\in \mathbb{Z}/M\mathbb{Z}
\end{equation}
where $\mathcal{S} \subset \widehat{\mathbb{G}} =\frac{1}{M}\mathbb{Z}/\mathbb{Z}$ is the set of
non-vanishing points of $\omega$ and $ (k,x) = \exp(2 \pi i kx) = \zeta_M^{Mkx} \in
\mathbb{Q}[\zeta_M]$, where $\zeta_M=\exp{2 \pi i /M}$ (the values of
$k$ are of the form $k_0/M \mod \mathbb{Z}$ where $k_0\in\mathbb{Z}$). Notice that for finite groups, the situation
we are in here, the densities $\rho_a$ are functions and also measures, since the distinction
between them disappears.
Gr\"unbaum and Moore \cite{GM} consider the case in which
$\rho_a$ is rational valued, i.e.
all of the weights $w_x$ are rational numbers. In this case, solving the system
of equations \eqref{weightEquations} for the `unknowns' $a(k) \omega(k)^{1/2}$
we see that they must all lie in $\mathbb{Q}[\zeta_M]$, and then
$a(k) \omega(k)^{1/2}\overline{a(k) \omega(k)^{1/2}} = \omega(k) \in \mathbb{Q}[\zeta_M]$ also.
Let $H:= {\rm Gal}(\mathbb{Q}[\zeta_M]/\mathbb{Q}) \simeq (\mathbb{Z}/M\mathbb{Z})^\times$, the group of units
of the ring $\mathbb{Z}/M\mathbb{Z}$. We can view $H$ as both a group of automorphisms
permuting the $M$ roots of unity, or as a group of automorphisms of $\frac{1}{M}\mathbb{Z}/\mathbb{Z}$;
namely $s\in (\mathbb{Z}/M\mathbb{Z})^\times$ acts as $\alpha_s:\exp(2 \pi i k) \mapsto \exp(2 \pi i sk)$
and as $\alpha_s:k \mapsto sk$ respectively. In the rational case we are now in, it is shown in
\cite{GM} that
$\alpha(\mathcal{S}) = \mathcal{S}$ for all $\alpha \in H$. We show next why this happens:
We take $a(k)\omega(k)^{1/2} = 0$ if $k\notin\mathcal{S}$.
If $k_1 \in \widehat{\mathbb{G}}$ then
\begin{equation}\label{orthogonalSum}
\frac{1}{|\mathbb{G}|} \sum_{x\in \widehat{\mathbb{G}}} k_1(x) \rho_a(x) =
\sum_{k\in \mathcal{S}} a(k) \omega(k)^{1/2} \frac{1}{|\mathbb{G}|} \sum_{x\in \widehat{\mathbb{G}}} k_1(x) \overline{k(x)} = a(k_1) \omega(k_1)^{1/2}\, ,
\end{equation}
which is zero if $k_1\notin\mathcal{S}$.
\smallskip
Let $\alpha = \alpha_s \in H$. Since $\rho_a$ is rational valued,
$\rho_a = \alpha(\rho_a)
= \sum_{k\in \mathcal{S}} \alpha(a(k) \omega(k)^{1/2}) \alpha(\overline k)$.
Applying the process of summation of equation \eqref{orthogonalSum} to the element $\alpha(k_1)$ where $k_1 \in \mathcal{S}$ we obtain $a(\alpha(k_1))\omega(\alpha(k_1))$. But applying $\alpha$ dircectly
to \eqref{orthogonalSum} we obtain
$\frac{1}{|\mathbb{G}|} \sum_{x\in \widehat{\mathbb{G}}} \alpha(k_1(x)) \alpha(\rho_a(x)) = \alpha(a(k_1)\omega(k_1)^{1/2})$. These two are equal, and this shows, in particular, that
$a(\alpha(k_1))\omega(\alpha(k_1))^{1/2} \ne 0$, and hence that $\alpha(k_1) \in \mathcal{S}$,
which is what we wished to prove.
In short, if we identify
$\mathcal{S}$ as a subset of $\frac{1}{M}\mathbb{Z}/\mathbb{Z}$, then $k \in \mathcal{S}$ and $s\in (\mathbb{Z}/M\mathbb{Z})^\times$
implies $sk \in \mathcal{S}$. In effect $\mathcal{S}$ is a union of orbits of $\frac{1}{M}\mathbb{Z}/\mathbb{Z}$
under the action of the multiplicative group $(\mathbb{Z}/M\mathbb{Z})^\times$.
This condition
appears in an equivalent form in the following key result of \cite{GM}. For
$h\in \mathbb{Z}$, let $\mbox{\rm ord}(h)$ denote its order as a group element modulo $M$
(under addition).
\begin{prop} \label{GMThm}If the density $\rho_a$ of \eqref{weightEquations} is rational
valued then for all $k\in \mathcal{S}$ and for all $j \in \mathbb{Z}$ with $\gcd(j, \mbox{\rm ord}(Mk)) =1$
one has $jk \in \mathcal{S}$. In this case any mapping $a:\mathcal{S} \longrightarrow U(1)$
satisfying the $m$-moment condition for all $m \le 6$ if $M$ is even (respectively
$m\le 4$ if $M$ is odd) extends to a character on $\frac{1}{M}\mathbb{Z}/\mathbb{Z}$. \qed
\end{prop}
Using Proposition~\ref{charChar}, and Corollary~\ref{finite-coro} and the
corresponding notation, we obtain
\begin{coro} Under the conditions of Proposition~\ref{GMThm}, the
set of distributions $\rho_a$ whose diffraction is $\omega$ of \eqref{finiteDiffraction}
is classified by the first $6$ moments (resp. first $4$ moments) if $M$ is even
(respectively if $M$ is odd). \qed
\end{coro}
We do not know of any extensions of this result into higher dimensional periodic
situations, and indeed it seems very difficult to do.
\begin{example} Here we give an example based on an example given in \cite{GM}
(see also \cite{XR3}).
We begin with $\mathbb{G}:= \mathbb{Z}/6\mathbb{Z}$ and the two weighted Dirac combs
\begin{eqnarray}\label{Z6 density}
11 \delta_0 + 25 \delta_1 + 42\delta_2 +45 \delta_3 + 31 \delta_4 + 14\delta_5\\
10 \delta_0 + 17 \delta_1 + 35 \delta_2 + 46 \delta_3 + 39 \delta_4 + 21 \delta_5\nonumber
\end{eqnarray}
on $\mathbb{G}$, which represent two density distributions for which we would like to find all possible
homometric density distributions. What is remarkable is that not only are these two density
distributions homometric, they actually have all of their first $5$ moments in common.
Let us begin with the first of these two densities. The Fourier transform of this density is a function on $\widehat{\mathbb{G}} \simeq \frac{1}{6}\mathbb{Z}/\mathbb{Z}$. The latter is most conveniently thought of as the $6$th roots of unity.
We write the Fourier transform in the form
\begin{eqnarray}
(5 + 25 w + 42 w^2&+&45 w^3 + 31 w^4 + 14 w^5 + 6 w^6)\\
&=&
(w + 1)(w^2 + w + 1)(2 w^2 + 5)(3 w + 1) =:P(w) \nonumber
\end{eqnarray}
where $w$ varies over the $6$th roots of $1$,
$U_6:= \{e^{2 \pi ij/6}: j=0, \dots, 5\}$. Note that $w^6= 1$, so the coefficient
of $w^0$ is $11/6$ as it is supposed to be. The diffraction at $k \in \widehat{\mathbb{G}}$ is $\omega(k) = P(w^k)P(\overline{w}^k)$, which works out
to give the following values for $\omega(k)^{1/2}:$ $ 28$, $ \sqrt{247/3}$, $ 0$ ,$0$ ,$0$ ,$ \sqrt{247/3}$
at $k=0,1/6,\dots, 5/6$.
Thus in this example $\mathcal{S} = \{0,1/6,5/6\}$. In fact Gr\"unbaum and Moore constructed
the polynomial $P$ precisely to make the Bragg spectrum have extinctions at
$2/6,3/6,4/6$. In the rest of this subsection we shall make the notation a little
simpler by suppressing the ubiquitous denominators
$6$ which occur in the values of $k$. Alternatively we multiply everything by
$6$ and work in $\mathbb{Z}/6\mathbb{Z}$ instead of $\frac{1}{6}\mathbb{Z}/\mathbb{Z}$.
In any $m$-tuple ${\bf k} = (k_1,\dots, k_m)$ in $\mathcal{S}^m$, we can, using the equivalence
relation $\sim$, drop any occurrences of $0$ and drop any pair $1,5$ that occurs amongst
the $k_j$. This leads us to conclude the only equivalence classes of $\mathbb{S}$ are of the form
$0^\sim, (1,\dots, 1)^\sim, (5, \dots, 5)^\sim$, where there are any positive integer
$m$ entries in the two latter types. Since $(1,\dots, 1)^\sim$ and $(5, \dots, 5)^\sim$
with $m$ entries are inverses of each other in $\mathbb{S}$, we conclude that
$\mathbb{S}^\sim \simeq \mathbb{Z}$. Since the mapping $\phi: \mathbb{S}^\sim\longrightarrow \mathcal{E}_d \simeq
U_6$, we have $\varphi:\mathcal{F}(\mathcal{S}) \longrightarrow U_6$ (Prop.~\ref{F(S) homomorphism}).
Using the identification of $\mathbb{S}^\sim$ and $\mathcal{F}(\mathcal{S})$ we have $\mathcal{F}(\mathcal{S}) \simeq \mathbb{Z}$
and we can put together the exact sequence \eqref{exactSeq1}:
\[1 \longrightarrow 6\mathbb{Z} \simeq \mathcal{Z} \longrightarrow \mathcal{F}(\mathcal{S}) \simeq \mathbb{Z}
\longrightarrow U_6 \simeq \mathcal{E}_d = \widehat{\mathbb{G}} \longrightarrow 1.\]
Dualization then gives the exact sequence \eqref{exactSeq2}
\[1 \longleftarrow \widehat{\mathcal{Z}} \simeq U(1)/U_6\simeq \frac{1}{6}\mathbb{Z}/\mathbb{Z} \longleftarrow \widehat{\mathcal{F}(\mathcal{S}) }\simeq U(1)
\longleftarrow \mathbb{Z}/6\mathbb{Z} \simeq \mathbb{T} \longleftarrow 1 .\]
Each element of $U(1)$ can be viewed as a character of $\frac{1}{6}\mathbb{Z}$,
and when restricted to $\mathbb{Z}$
characters $a(\cdot)$ and $e^{2\pi ij(\cdot)} a(\cdot)$, $j=0,\dots, 5/6$, are the same.
We now write down the densities
\[\rho_a(x) = \sum_{k\in \widehat{\mathbb{G}}} a(k) \omega(k)^{1/2} \overline{k}(x) =
28 + \sqrt{247/3} \, a(1) e^{-2\pi i x/6}+ \sqrt{247/3} \,a(5) e^{2 \pi i x/6} \,.\]
Here $x\in \mathbb{G} \simeq \mathbb{Z}/6\mathbb{Z}$. The character $a$ of
$\mathcal{F}(\mathcal{S})$ is a character of $\mathbb{Z}$ and hence is an element of $U(1)$. Recall
that $\mathcal{F}(\mathcal{S})$ is generated by the tuples $(1,1,\dots)$ and their inverses
$(5,5,\dots,5)$ where $1,5$ represent classes modulo $6$ in $\mathbb{Z}$. There are
no further relations. Thus $a(k)$ here really means $a((k)^\sim)$, that is, the value
of $a$ on the the equivalence class of $(k)$ under $\sim$. Thus
by $a(1)$ we mean $a((1)^\sim)$ and by $a(5)$ we mean
$a((5)^\sim) = \overline{a(1)}$. There is no restriction on what
$a(1)$ might be in $U(1)$, so simply writing it as $a$ we may write $\rho$ as
\[\rho_a(x) =
28 + \sqrt{247/3} \, a e^{-2\pi i x/6}+ \sqrt{247/3} \,\overline a e^{2 \pi i x/6} \,\]
where $a\in U(1)$ is arbitrary. Each value of $a$ gives a solution to the
diffraction problem.
The following figure illustrates the graphs of $\rho_a(x)$.
The situation that we began with occurs when $a \sim 0.443099$.
Another solution indicated in \cite{GM} is $10, 17, 35, 46, 39, 21$ which occurs
at $a\sim 0.520310 $.
\begin{figure}
\centering
\includegraphics[width=7cm]{Z6DiffractionGraph.pdf}
\caption{The graphs of
$\rho_a(x)$ as $a$ varies for each of the six values $0,1,2,3,4,5 \mod 6$ of $x$.
Here $a$ varies over the unit circle $\{e^{2 \pi i t}: t\in [0,1)\}$, indicated here by the interval $[0,1)$
and the six graphs give the corresponding values of the coefficients or weights of the deltas
that make up the distribution on $\mathbb{Z}/6\mathbb{Z}$. The original weighting distribution
$11, 25 ,42, 45 ,31 ,14$ occurs when $a \sim 0.443099$. The second
density of \eqref{Z6 density} with coefficients $10, 17, 35, 46, 39, 21$ also occurs,
when $a \sim 0.520310$. }
\label{Z6Graph}
\end{figure}
Now we are going to show that in this example, $\langle \mathcal{Z}_6 \rangle = \mathcal{Z}$,
but $\langle \mathcal{Z}_5 \rangle \subsetneq \mathcal{Z}$. The argument is quite simple. First note
that $(1,1,1,1,1,1) \in \mathcal{Z}_6$ but cannot be reduced, so its reduced length is $6$ (note
that denominators are still being suppressed).
So $\langle \mathcal{Z}_5 \rangle \ne \mathcal{Z}$. Now consider any ${\bf k} = (k_1,\dots, k_n)\in Z_n$
with $n \ge 6$, which is of minimal length and can not be split into a product of two (non trivial) elements from $\mathcal{Z}$. There can be no occurrences of $0$ since those can be discarded. Since $\{1,5\}$ and $\{2,4\}$ are annihilating pairs modulo $6$, we cannot
have both $1$ and $5$ in ${\bf k}$ and neither can we have both $2$ and $4$. Neither
can we have $2$ appearing three or more times since $(2,2,2) \in Z_3$ and similarly
for $4$. Likewise $3$ can appear at most once. Suppose $2$ or $3$ appears in ${\bf k}$.
The possibilities are $2,22, 3, 23, 223$. The remaining symbols are just $1$s or
$5$s. But $21111, 2211, 3111,255, 3555$ then show that ${\bf k}$ is not reduced. The
argument with $4$ is much the same. So we can assume that none of $0,2,3,4$ appear
and not both $1$ and $5$ appear. This leaves just $(1,\dots,1)$ and $(5,\dots,5)$
as the possibilities for reduced words, and in either case the number of entries must be
a multiple of $6$.
The consequence of this is that any phase form is uniquely determined by its
first through sixth moments. For if two have these moments in common then
taking corresponding elementary phase forms $a,b$ we find that $ba^{-1} \in \mathcal{F}(\mathcal{S})$
and satisfies the first $6$ moment conditions. Since $\langle \mathcal{Z}_6 \rangle = \mathcal{Z}$, $ba^{-1}$
lies in $\mathbb{T}$ and so the corresponding phase forms were equal.
It is a surprising fact, pointed out in \cite{GM}, that the two densities of \eqref{Z6 density},
$11 \delta_0 + 25 \delta_1 + 42\delta_2 +45 \delta_3 + 31 \delta_4 + 14\delta_5$ and
$10 \delta_0 + 17 \delta_1 + 35 \delta_2 + 46 \delta_3 + 39 \delta_4 + 21 \delta_5$
indicated above, have all of their first through fifth moments in common, and it is
only at the sixth moments that they look different.
\end{example}
\section{An alternative construction of spatial processes from pure point diffraction}\label{altConstruction}
In \S\ref{CyclestoStochasticprocesses} we offered a direct method of constructing a spatial process from the original data consisting of a positive backward transformable point measure
$\omega$ on $\widehat{\GG}$ and a phase form $a^*$ (or equivalently an elementary
phase form $a$). This approach can in a certain sense be understood as an elaboration of the theorem of Halmos - von Neumann which
tells us in advance that since the required process will have an associated
dynamical system with pure point spectrum, it can be modelled (measure theoretically) on
a compact Abelian group.
In \S\ref{pc:split} we have used the Gelfand method to produce pure point spatial processes in the splitting of spatial processes into their pure and continuous parts.
In this present section we sketch out how we can also use the Gelfand method to produce a pure point spatial process out of the given pure point measure $\omega$ and the elementary phase form $a$. The method ultimately reverses many of the steps that we saw earlier in which, starting
from a process we constructed the diffraction and the diffraction-to-dynamics map. For this reason we do not give proofs but rather provide an outline of intermediate steps.
\bigskip
We use the notation of \S\ref{Cycles}. For simplicity of presentation we shall assume
that $0\in \mathcal{S}$.
Define $c :\mathcal{S} \longrightarrow \mathbb{C}$ by
\[ c(k_1, \dots, k_n) = a(k_1, \dots, k_n )\omega(k_1)^\frac{1}{2}\dots \omega(k_n)^\frac{1}{2} \, , \]
or more compactly,
$c({\bf k}) = a({\bf k})\omega({\bf k})^\frac{1}{2}$,
where for each square root we take the non-negative root. We define
$c(\emptyset) = 1$. Note that $c({\bf k}({-\bf k})) = \omega({\bf k})$.
Let
$\mathbb{C}[F_k : k \in \mathcal{S}]$
be the free associative algebra in
the variables $F_k$. For elements ${\bf k} = (k_1, \dots , k_m) $
where the $k_j \in \mathcal{S}$, we will write
$F_{\bf k} := F_{k_1} \dots F_{k_m}$.
We define an
algebra $\mathcal{E}$-grading on $\mathbb{C}[F_k : k \in \mathcal{S}]$ by assigning degree $k$ to
each $F_k$, so $F_{{\bf{k}}}$ gets degree $[{\bf{k}}]$.
Let $I = I(a)$ denote the ideal of $\mathbb{C}[F_k : k \in \mathcal{S}]$ generated by
all the elements of the form
$F_{\bf k} - c({\bf k})$
for all ${\bf k} = (k_1, \dots k_m) \in Z$, i.e. for
all ${\bf k}$ with $[{\bf k}] = 0$ (this includes
$F_0 - \omega(0)^{1/2} 1,\quad F_k F_{-k} - \omega(k)$)
and let
\[ \mathcal{P} = \mathcal{P}(a) := \mathbb{C}[F_k : k \in \mathcal{S}]/I \,. \]
We let \[\alpha: \mathbb{C}[F_k : k \in \mathcal{S}] \longrightarrow \mathcal{P}\]
be the natural homomorphism and set
\[ f_k := \alpha(F_k), \quad f_{{\bf{k}}} := \alpha(F_{{\bf{k}}}) = f_{k_1} \dots f_{k_m} \]
for ${\bf{k}} = k_1 + \dots + k_m$.
Since $c(k_1, \dots, k_m)$ is independent of the order of the $k_j$,
we see that $\mathcal{P}$ is commutative.
Furthermore, since the relations imposed on $\mathbb{C}[f_k : k \in \mathcal{S}]$ are homogeneous of degree $0$, the resulting algebra $\mathcal{P}(a)$ is also graded by $\mathcal{E}$. We denote its space of elements of degree $\kappa \in \mathcal{E}$ by $\mathcal{P}_\kappa$.
The space $\mathbb{C}[f_k : k \in \mathcal{S}]_\kappa$ of elements of degree $\kappa \in \mathcal{E}$
is the linear span of all the elements $f_{{\bf{k}}}$ with $[{\bf{k}}] = \kappa$. If
$[{\bf{k}}] = [{\bf{l}}] = \kappa$ then ${\bf{k}}(-{\bf{l}}) \in Z$ and the relations defining $I$
lead to
\begin{equation} \label{invertibility}
f_{{\bf{k}}} f_{-{\bf{l}}} = f_{k_1} \dots f_{k_m} f_{-l_1} \dots f_{-l_n}
=a({\bf{k}} (-{\bf{l}})) \omega({\bf{k}})^{1/2} \omega(-{\bf{l}})^{1/2} \in \mathbb{C}^\times 1 \,.
\end{equation}
From this we see first that
$ \mathcal{P}_\kappa = \mathbb{C} f_\kappa$
where $f_\kappa$ is any one of the elements $f_{{\bf{k}}}$
with $[{\bf{k}}] =\kappa$.
We shall always assume that we are working with a phase form $a$ for which $\mathcal{P}(a)$ is not reduced to the trivial ring $\{0\}$.
Such cycles always exist, for example the trivial cycle, which is $1$ everywhere.
Since the algebra is $\mathcal{E}$-graded, non-zero elements of different degrees are linearly independent
and
\begin{equation*}
\mathcal{P} = \mathcal{P}(a) = \bigoplus_{\kappa \in \mathcal{E}} \mathbb{C} f_\kappa \, .
\end{equation*}
Define an conjugate linear form on $\mathbb{C}[f_k : k \in \mathcal{S}]$ by sesqui-linear extension (with conjugation on the second variable) of
\[\langle f_{\bf k} \mid f_{\bf l} \rangle =
\langle f_{k_1} \dots f_{k_m} \mid f_{l_1} \dots f_{l_n}\rangle
= \left \{ \begin{array}{ll}
c(({\bf k}) (-{\bf l})) \; \mbox{if \; $[{\bf k}] = [{\bf l}]$} \, ,\\
0 \quad \mbox{otherwise} \, .
\end{array}
\right.
\]
It is a straightforward exercise to show the identity
\[ \langle f \mid g \rangle = \overline{\langle g \mid f \rangle} \]
for all $f,g \in \mathbb{C}[f_k : k \in \mathcal{S}]$. We shall write $||f||$ for $\langle f \mid f \rangle^{1/2}$.
\begin{prop} Let ${\bf k}, {\bf l}, {\bf m}$ be tuples of elements of
$\mathcal{S}$.
\begin{itemize}
\item[{(i)}] $\langle f_{\bf k} \mid f_{\bf k}\rangle = c({\bf k}(-{\bf k})) = \omega({\bf k}) >0$;
\item[{(ii)}] $\langle 1 \mid 1\rangle = 1$; $\langle f_{\bf{l}}\,| \,1 \rangle = c({\bf{l}})$
\mbox{ when} $ [\bf l] = 0$;
\item[{(iii)}] $\langle f_{\bf k} f_{\bf m} \mid f_{\bf l} \rangle
= \langle f_{\bf k} \mid f_{-\bf m}f_{\bf l} \rangle$;
\item[{(iv)}]
$\langle f_{\bf k} - c({\bf k})1 \mid f_{\bf l} \rangle = 0 $ \; when $[{\bf{k}}] =0$;
\item [{(v)}] The kernel of $\langle \cdot \mid \cdot \rangle $ is the ideal of relations
that we have imposed on $\mathbb{C}[f_k : k \in \mathcal{S}]$;
\item[{(vi)}] $\langle \cdot \mid \cdot \rangle $ induces an inner product (for which we use
the same notation) on $\mathcal{P}$.
\end{itemize}
\end{prop}
Now $\mathcal{P}$ is an inner product space. We wish next to consider $\mathcal{P}$
acting on itself by multiplication and to make it into a normed algebra
by use of the operator norm. The details are as follows.
A typical element of $\mathcal{P}$ can be written as a finite sum
$x = \sum b_{\bf{l}} f_{\bf{l}}$ with $b_{\bf{l}} \in \mathbb{C}$, where there is only one
summand of any particular degree (i.e. $[{\bf{l}}] \ne [{\bf{l}}'] $ if
${\bf{l}} \ne {\bf{l}}' $).
Then we have
\[ ||x||^2 =\sum |b_{\bf{l}}|^2 ||f_{\bf{l}}||^2 = \sum |b_{\bf{l}} |^2 \omega({\bf{l}}) \,.\]
Let ${\bf{k}} = (k_1, \dots, k_m) \in \mathcal{S}^m$. Then $f_{\bf{k}} x = \sum b_{\bf{l}} f_{\bf{k}} f_{\bf{l}} = \sum b_{\bf{l}} f_{{\bf{k}}{\bf{l}}}$ with degrees
$[{\bf{k}}] + [{\bf{l}}]$ and
\begin{eqnarray*}
||f_{\bf{k}} x||^2 &=& \langle \sum b_{\bf{l}} f_{{\bf{k}}{\bf{l}}} \mid \sum b_{\bf{l}} f_{{\bf{k}}{\bf{l}}}\rangle^{1/2}\\
&=& \sum |b_{\bf{l}}|^2 ||f_{{\bf{k}} {\bf{l}}}||^2 = \sum |b_{\bf{l}}|^2 \omega({\bf{k}} {\bf{l}})
=\omega({\bf{k}} ) \sum |b_{\bf{l}}|^2 \omega({\bf{l}})|| = ||f_{\bf{k}}||^2||x||^2\,.
\end{eqnarray*}
Thus
\begin{equation} \label{norm}
||f_{\bf{k}} x|| = ||f_{\bf{k}}|| \,||x|| \,
\end{equation}
which shows that for each tuple ${\bf{k}}$, left multiplication by $f_{\bf{k}}$ is a linear operator of norm $||f_{\bf{k}}||$.
Now we see that the entire left representation
of $\mathcal{P}$ on itself is a representation by bounded linear operators, and this provides
us with an operator norm $\nu$ on $\mathcal{P}$ with $\nu(f_{\bf{k}}) = ||f_{\bf{k}}||$ and
$\nu(\sum b_{\bf{l}} f_{\bf{l}}) \le \sum |b_{\bf{l}}| \nu(f_{\bf{l}})$ in general. In view of \eqref{norm}
we always have
\[ \nu(f_{\bf{k}}) = \omega(\{{\bf{k}}\})^{1/2} \;\mbox{and} \; ||x|| \le \nu(x)\; \mbox{ for all } \; x \in \mathcal{P} \,.\]
Being an operator norm, $\nu$ automatically satisfies
\[\nu(xy) \le \nu(x) \, \nu(y) \,.\]
\smallskip
We let $\mathcal{A}$ denote the completion of $\mathcal{P}$ with respect to $\nu$. This is a commutative Banach algebra. The norm will still be called $\nu$. We identify
$\mathcal{P}$ with its image in $\mathcal{A}$. Since $\nu(f) \ge ||f||$ for $f\in \mathcal{P}$, the embedding
of $\mathcal{P}$ in $\mathcal{A}$ is faithful.
We define an $\mathbb{G}$ action on $\mathcal{P}(a)$ by
\[ T_t f_{k_1} \dots f_{k_m} = (k_1 + \dots + k_m)(t) f_{k_1} \dots f_{k_m} \, ,\]
or in abbreviated form
$T_t f_{\bf{k}} = [{\bf{k}}](t) f_{\bf{k}}$,
for all $t\in \mathbb{G}$. We note that $T_t$ acts as a unitary transformation relative
to the inner product and
we obtain a unitary representation $T$ of $\mathbb{G}$ on $\mathcal{P}$.
Each $T_t$ is an automorphism of $\mathcal{A}$ as a Banach
algebra.
At this point we can apply Gelfand theory to produce the space $X = {\rm Spec}(\mathcal{A})$.
which is the set of algebra
homomorphisms $\pi: \mathcal{A} \longrightarrow \mathbb{C}$.
Then $X$ is non-empty and compact in the Gelfand topology \cite{Lax}.
This is the weakest topology on $X$ making continuous the evaluation mappings
$ \eta_f: \pi \mapsto \pi(f)$
for each $f\in \mathcal{A}$. This means that the open sets are generated
from the open sets
\[U(f,V) := \{ \pi \in X: \pi(f) \in V \}\, ,\]
where $f \in \mathcal{A}$, $V$ open in $\mathbb{C}$.
This is the way in which $\mathcal{A}$ may be viewed as a space of continuous mappings on
$X$, namely $f(\pi) := \pi(f)$, and this is the point of view that we take from now on.
One has for all $\pi \in X$, $f\in \mathcal{A}$,
$ |\pi(f)| \le \nu(f)$
(see \cite{Lax}, Ch. 18, Prop.~1) from which for the sup norm we have
$||f||_\infty \le \nu(f)$.
For all tuples ${\bf{k}}$ from $\mathcal{S}$ and for all $\pi \in X$,
\[ |f_{\bf{k}} (\pi)| = ||f_{\bf{k}}|| \,.\]
Since $T_t$ acts as an algebra automorphism of $\mathcal{A}$, $T_t \pi =\pi \circ T_{-t}$
is another homomorphism of $\mathcal{A}$ into $\mathbb{C}$, and so is in $X$.
So $\mathbb{G}$ acts on $X$. Since $T_t$ just permutes the defining open sets of the topology of $X$, this action is continuous.
\begin{lemma} \label{continuity}
For each $g\in \mathcal{A}$ the mapping
\[ \mathbb{G} \longrightarrow \mathbb{C}, \quad \quad t \mapsto \nu(T_t g - g) \]
is uniformly continuous.
\end{lemma}
\begin{prop} \label{bicontinuity}
The mapping
\[ \mathbb{G} \times X \longrightarrow X \]
defined by the action of $\mathbb{G}$ on $X$ is continuous.
$(X,\mathbb{G})$ is a topological dynamical system.
\end{prop}
Since $\mathbb{G}$ acts on $X$, it acts naturally on $C(X)$ too. $\mathcal{P}$ already has an action of $\mathbb{G}$ and acquires another one when treated as a subalgbra of $C(X)$; however, not surprisingly, the two actions of $\mathbb{G}$ on $\mathcal{P}$ the same.
\begin{prop} $\mathcal{P}$ is dense in the space $C(X)$ of all continuous functions under the sup-norm.
\end{prop}
\medskip
The next objective is to create an ergodic invariant measure for $(X,\mathbb{G})$.
We define a linear functional $\mu = \mu_{a}$ on $\mathcal{P}$ by linear extension of
\begin{equation*}
\mu( f_{\bf{k}}) :=
\begin{cases} c({\bf{k}}), &\text{if $[{\bf{k}}] = 0$}\,;\\
0 &\text{otherwise}
\end{cases}\end{equation*}
for all tuples ${\bf{k}}$ from $\mathcal{E}$. Note that this is well-defined -- the relations amongst
the $f_{\bf{k}}$ for equal values of $[{\bf{k}}]$ are consistent with the definition --
and one can see immediately that $\mu$ is $\mathbb{G}$-invariant since it picks up only the
$0$-eigenspace of $\mathcal{P}$.
\smallskip
We need to extend this to a linear functional defined on all of $C(X)$.
Let $\{A_n\}_{n=1}^\infty$ be a van Hove sequence for $\mathbb{G}$. For each
$f \in C(X)$ define a new function $M(f)$ on $X$ by
\[ M(f)(\pi) := \lim_{n\to\infty} \frac{1}{\vol A_n} \int_{A_n} (T_t f)(\pi) \, dt \,, \]
where integration is with some prefixed Haar measure on $\mathbb{G}$ and
$\vol(A_n)$ is the value of this measure on $A_n$.
\begin{lemma} \label{mean1}
For all $f \in \mathcal{P}$, $M(f)$ is a constant function. Moreover,
\[M(f_{\bf{k}}) = \begin{cases}
c({{\bf{k}}}), &\text{if \; $[{\bf{k}}] = 0$} \,;\\
0 \ &\text{otherwise} \,.
\end{cases}\]
\end{lemma}
This shows that $\mu$ is the same as the mean $M$ on $\mathcal{P}$. But, as we now see,
$M$ is defined on all of $C(X)$ and is a bounded linear operator on $C(X)$:
\begin{lemma} \label{mean2}
For all $g \in C(X)$, $||M(g)||_\infty \le ||g||_\infty$. For all $g \in C(X)$,
$M(g)$ is a constant function.
\end{lemma}
In this way, $M$ produces a measure on $X$, which coincides
on $\mathcal{P}$ with $\mu$ defined above. We also call this measure $\mu$ and note
the fact that we have established, namely for all $g\in C(X)$, and for all $\pi \in X$,
\[ \mu(g) =\lim_{n\to\infty} \frac{1}{\vol A_n} \int_{A_n} (T_t g)(\pi) \, dt \, , \]
which is a form of ergodic theorem.
\begin{prop} \label{cc}
For each $\kappa \in \mathcal{E}$, the space of $\kappa$-eigenfuctions
in $C(X)$ (for the action of $T$ on $X$) is one dimensional and it is spanned by
any of the functions $f_{\bf{k}}$ for which $[{\bf{k}}] = \kappa$. In particular the $0$-eigenfunctions are constant functions. The complex conjugate of $f_{\bf{k}}$
as a function on $X$ is $f_{-{\bf{k}}} = c({\bf{k}}(-{\bf{k}}))f_{\bf{k}}^{-1}$.
\end{prop}
\begin{prop} \label{mu=measure}
$\mu$ is $\mathbb{G}$-invariant probability measure on $X$ and
$(X,\mathbb{G},\mu)$ is a pure point ergodic dynamical system.
\end{prop}
\medskip
We now wish to define the ergodic spatial process $N$. To do this
we first define the diffraction-to-dynamics map.
For each $k \in \mathcal{S}$, let $1_k : \widehat{\mathbb{G}} \longrightarrow \mathbb{C}$
be the function which is $1$ at $k$ and $0$ everywhere else.
Since the measure $\omega$ is only non-zero on the points of $\mathcal{S}$,
it follows that every function in $L^2(\widehat{\mathbb{G}}, \omega)$ can be represented
in the form
\[ \sum_{k \in \mathcal{S}} x_k \,1_k, \; \mbox{where} \;
\sum_{k \in \mathcal{S}} |x_k|^2 \omega(k) <\infty \, . \]
The diffraction-to-dynamics embedding is the mapping
\begin{eqnarray*}
\theta = \theta_{\tilde a}\,:\,L^2(\widehat{\mathbb{G}}, \omega) &\longrightarrow& L^2(X,\mu)\\
\sum_{k \in \mathcal{S}} x_k 1_k &\mapsto& \sum_{k \in \mathcal{S}} x_k f_k \,.
\end{eqnarray*}
Comparing these two $L^2$-spaces we have
\[ \langle \sum x_k 1_k \, \mid \sum x_k 1_k\rangle =
\int_{\widehat{\mathbb{G}}} \sum_{k,l} x_k \overline{x_l }\, 1_k \, 1_l d \,\omega=
\sum_k |x_k|^2 \omega(k)
= \langle \sum x_k f_k \mid \sum x_k f_k \rangle \, ,\]
which shows that $\theta$ is an isometric embedding.
Define an action $U$ of $\mathbb{G}$ on $L^2(\widehat{\mathbb{G}}, \omega)$ by setting, for each
$g\in L^2(\widehat{\mathbb{G}},\omega)$, $U_t(g)$ to be the function
\begin{equation*}
U_t(g)(k) = k(t)g(k) \quad \mbox{for all} \, k\in \widehat{\mathbb{G}} \, .
\end{equation*}
Under this action, $1_k \in L^2(\widehat{\mathbb{G}}, \omega)$ is a $k$-eigenfunction for the action
and $\theta$ becomes $\mathbb{G}$-equivariant.
Since $\omega$ is a positive, translation bounded, and backward transformable measure on $\widehat{\mathbb{G}}$ there is a
unique positive definite measure $\gamma$ on $\mathbb{G}$ for which $\omega
= \hat\gamma$. This measure is also translation bounded (and hence
automatically transformable (with Fourier transform equal to $\omega$), \cite{BF}). The basic relationship between $\gamma$ and $\omega$ is expressed by \eqref{omega-gamma}
for all $F \in C_c(\mathbb{G})$, see \cite{BF}, Ch.~1.4, which in particular gives $\hat F \in L^2(\widehat{\mathbb{G}}, \omega)$. Direct calculation shows that for $F\in C_c(\mathbb{G})$,
$U_t \hat F = \widehat{T_{-t} F}$.
We define the process by
\[ N=N_a : C_c(\mathbb{G}) \longrightarrow L^2(X, \mu), \quad N(F) = \theta(\hat F) \, .\]
Also $N$ is {\bf real}, linear, continuous, real, and stationary ($\mathbb{G}$-equivariant). It is
also full.
\begin{prop} \label{stochasticProcess}
$\mathcal{N} = (N,X,\mu,T)$ is a pure point ergodic spatial process with diffraction $\omega$
and associated phase form $a^*$.
\end{prop}
Thus, based on the positive measure $\omega$ on $\widehat{\mathbb{G}}$
and a phase form $a^*$ we have constructed a pure point spacial process on $\mathbb{G}$.
\section{Various concluding remarks}\label{conclusion}
In this section we collect together various small items that are of interest.
\subsection{The case that $N$ maps into $L^\infty$}\label{Special}
A spatial stationary process $\mathcal{N}= (N,X,\mu,T)$ is said to have an $m$th moment if
$N(F_1)\ldots N (F_m)$ belongs to $L^1 (X,\mu)$ for any $F_1,\ldots, F_m\in C_c (\mathbb{G})$. In this case, the $m$th moment is defined to be the unique linear map $\mu^{(m)}$ on $C_c (\mathbb{G}) \otimes \cdots \otimes C_c (\mathbb{G})$ ($m$ factors) with
$$\mu^{(m)} (F_1\otimes \cdots \otimes F_m) := \int N(F_1)\cdots N(F_m) d\mu. $$
It it not hard to see that the measure $\mu$ determined by its moments $\mu^{(m)}$, $m\in \mathbb{N}$, if all these moments exist and the process if full. In fact, this is more or less the definition of fullness. In order to have a meaningful diffraction
theory our running assumption is that the second moment exists and is a measure. Then, also the first moment is a measure (as can easily be seen).
We say that $\mathcal{N}$ is {\it bounded} if $N(F)\in L^\infty (X,\mu)$ for every $F\in C_c (\mathbb{G})$. In this case, for every $m\in \mathbb{N}$ the $m$th moment exists.
We are now going to discuss how for bounded processes this notion of moment is related to the notion of moment of a phase form introduced above.
Thus, let $(N, X,\mu,T)$ be an bounded, ergodic full stationary process with pure point spectrum and elementary phase form $a$. Then, the $m$th moment of $\mu$ and the $m$-moment of $a$ uniquely determine each other. More precisely, the following holds in this situation.
\begin{lemma} (a) Let $n\in \mathbb{N}$ and $F_1,\ldots, F_n \in C_c (\mathbb{G})$ be given. Then,
$$\int N ( F_1) \dots N (F_n) d\mu = \sum_{k_1\in
\mathcal{S}}\ldots\sum_{k_n\in \mathcal{S}} \widehat{F_1} (k_1)\dots \widehat{F_n}
(k_n) \int_X (f_{k_1}\dots f_{k_n}) d\mu$$
where the sums exist if taken one after the other, and
for any $k_1, \dots, k_n \in \mathcal{S}$,
$$
\int_X f_{k_1} \dots f_{k_m} d\mu
= \begin{cases} a(k_1, \dots, k_m) \omega(k_1)^{1/2} \dots \omega(k_m)^{1/2} &\mbox{if} \; k_1 + \cdots +k_m = 0;\\
0 &\mbox{if} \; k_1 + \cdots +k_m \ne 0 \,.
\end{cases}
$$
(b) For
each $j=1,\ldots,n$, let $\{F_j^{(m)}\}$ be a sequence in $C_c (\mathbb{G})$ whose
Fourier transforms converge to ${\bf 1}_{k_j}$ in $L^2 (\widehat{\mathbb{G}},\omega)$.
Then, $$f_{k_1}\dots f_{k_n} = \lim_{m_1\to \infty} \lim_{m_2\to \infty}
\ldots \lim_{m_n\to \infty} N(F_1^{(m_1)}) \dots N(F_n^{(m_n)}),$$
where the limits are taken in $L^2$ and one after the other.
In particular, $$\int_X f_{k_1}\dots f_{k_n} d\mu = \lim_{m_1\to \infty}
\lim_{m_2\to \infty} \ldots \lim_{m_n\to \infty} \int_X N (F_1^{(m_1)}) \dots N(F_n^{(m_n)}) d\mu \,.$$
\end{lemma}
\begin{proof} This is essentially proven in Lemma 2.1 and Lemma 2.3 of \cite{LM} for special uniformly discrete point processes. The proof given there carries over to our situation.
\end{proof}
\begin{coro}
Let two bounded spatial processes with pure point diffraction and the same diffraction measure $\omega$ be given. Then, these processes have the same $m$-moment if and only if their associated phase forms have the same $m$ moment. In particular, if furthermore $\langle \mathcal{Z}_n\rangle = \mathcal{Z}$, then the two processes are isomorphic if and only if their $m$ moments agree for $m= 1,\ldots, n$.
\end{coro}
If every element of $\mathcal{E}$ can be expressed as a sum of $m$ or less elements
of $\mathcal{S}$, then $\langle \mathcal{Z}_{2m+1}\rangle = \mathcal{Z}$, and so $2m+1$ moments
suffice to determine the isomorphism class of a process, see Remark~\ref{2m+1 theorem}.
\subsection{A topological situation.}
In this section we look at a pure point stationary process
$N: C_c(\mathbb{G}) \longrightarrow L^2(X,\mu)$ and assume that $X$ is a topological space and all eigenfunctions are continuous. Here, continuity means that
for each eigenvektor $k$, we have an continuous nontrivial function $g_k$ with
$$g_k (T_t x) = (k,t)g_k (x)$$
for all $x\in X$ and $t\in \mathbb{G}$. Without loss of generality we can then normalize the $g_k$ to satisfy
$$ g_k (x_0) = 1$$
for some fixed $x_0 \in X$. These normalized $g_k$ will then obey
\begin{equation}\label{character}
g_{k_1} g_{k_2} = g_{k_1 + k_2}\;\:\mbox{and}\:\: g_{-k_1} = \overline{g_{k_1}}.
\end{equation}
Of course, as discussed in Section \ref{StochasticprocessestoDiffraction}, the pure point process comes with a diffraction measure
$\omega$ a set of eigenvalues $\mathcal{E}$ and a phase form $a$. Set $\mathbb{T}:= \widehat{\mathcal{E}_d}$. Then the considerations of Section \ref{StochasticprocessestoDiffraction} provide a torus system $(N_a, \mathbb{T}, l_{\mathbb{T}}, T)$. This system can be directly calculated as follows:
\begin{theorem} Assume the situation outlined above. Then,
$$\pi : X\longrightarrow \mathbb{T}, x\mapsto (k \mapsto g_k (x)),$$
is a surjective $\mathbb{G}$-map (and hence $\mathbb{T}$ is a factor of $X$) and
$$ M : L^2 (\mathbb{T})\longrightarrow L^2 (X), g\mapsto g\circ \pi,$$
is an isomorphism of point processes.
\end{theorem}
\begin{proof}
Due to the normalization and \eqref{character}, $\pi (x)$ belongs indeed to $\mathbb{T}$. Obviously, $\pi$ is continuous (as $\mathcal{E}$ is given the discrete topology).
Moreover, $M$ can easily be seen to map the character $k\in \mathcal{E}$ of $\mathbb{T}$ to the eigenfunction $g_k$. Thus, $M$ is unitary (as it maps an orthonormal basis to an orthonormal basis).
As $M$ is an isometry, $\pi$ must have dense image. As it it a $\mathbb{G}$ map it is then onto. Moreover, a short calculation
invoking \eqref{character} again shows that
$$M (g_{k_1} g_{k_2}) = g_{k_1} g_{k_2}.$$
By usual limiting arguments, this gives that $M (f g) = M (f) M (g)$ for all $f,g\in L^\infty$.
As $M$ maps the character $k\in \mathcal{E}$ to the eigenfunction $g_k$, its inverse $M^{-1}$ must map the eigenfunction $g_k$ to the character $k$. Combined with the usual limiting arguments this shows that $M^{-1}$ also satisfies $M^{-1} (u v ) = M^{-1} (u) M^{-1} (v)$ for bounded $u,v$.
So, if we now define the $N$-function $N_{\mathbb{T}}$ on the torus by
$$N_{\mathbb{T}} := M^{-1} N$$
we have indeed an isomorphism of processes. Thus, the phase form belonging to $N_{\mathbb{T}}$ must then be given by $a$ as well. By uniqueness then $[N_a] = [N_{\mathbb{T}}]$.
\end{proof}
Another issue around continuity is the possible continuity of
the mappings
$a:\mathcal{S} \longrightarrow U(1)$ in the topology on $\mathcal{S}$
induced from $\widehat{\mathbb{G}}$. For these one uses this same topology to make
$\mathcal{F}(\mathcal{S})$ into a topological group. We do not know if continuity here is
physically relevant but we do note that
in this case the classification results for elementary phase forms
compared to actual phase forms change
from $\mathbb{T}$ to $\mathbb{G}$ (see the discussion around and including Remark~\ref{continuousCase}).
\subsection{Spatial processes arising from measures on $\mathbb{G}$}\label{ssp-measures}
Our underlying understanding of a spatial process $\mathcal{N}=(N,X,\mu,T)$ is that it
represents some sort of `density' on the space
$\mathbb{G}$. Each point $\xi \in X$ represents an instance of the yet unspecified
structure $\rho_\xi$ which can be paired with elements of $C_c (\mathbb{G})$ to give
\begin{equation}
\label{measureInterpretation2}
\langle \rho_\xi, F\rangle = N(F)(\xi)\,.
\end{equation}
Thus $C_c(\mathbb{G})$ provides a set of test functions for some sort of distribution $\rho_\xi$
at $\xi \in X$. Of course $N(F)$ is not necessarily defined at all points of $X$
(it is only an $L^2$-function on $X$), and even if it is defined it is not clear whether
or not $\rho_\xi$ could be interpreted as any particular kind of distribution.
We do note however, that from $\mathbb{G}$-invariance we have
$N(F)(T_t \rho_\xi) = (T_{-t} N(F))(\xi) = N(T_{-t} F)(\xi)$, so
\[\langle T_t \rho_\xi, F\rangle = \langle \rho_\xi, T_{-t}F\rangle\,\]
for all $t\in \mathbb{G}$.
An important question, and one for which we cannot give a satisfactory answer,
is when can this `density' be interpreted as a measure $\rho_\xi$ on $X = \mathbb{G}$?
In the theory of Delone point sets $\Lambda$ this actually happens; typically
$X$ is the orbit closure of $\Lambda$ under the local topology and $N$ is
the process that arises from the point measures $\delta_\xi :=
\sum_{x\in \xi} \delta_x$ for all $\xi \in X$. Thus for $F \in C_c(\mathbb{G})$,
$N(F)(\xi) := \sum_{x\in \Lambda} F(x) = \langle \delta_\xi, F\rangle$.
\medskip
We have already seen one way in which we can see from general principles that measures arise
in \S\ref{compactG}. But this
is unnecessarily restrictive since in many basic situations $\omega$ is not summable. For example,
in \S~\ref{G=U(1)} we looked at
the diffraction of the point set $\delta_\mathbb{Z}$, when it is
modeled on $U(1)$ as $\delta_0$. Its diffraction is $\omega = \delta_\mathbb{Z}$ on $\mathbb{Z}$.
This is not $L^1$ with respect to the Haar measure of $\mathbb{Z}$ (counting measure).
Here $X$ can be identified with $U(1)$ (since it is compact) and for all $t \in U(1)$,
$F\mapsto N(F)(t)$ is just the measure $\delta_t$. So in this case there is a suitable
measure in spite of the lack of summability.
The exact conditions that $\rho_\xi$ defined by \eqref{measureInterpretation2} should be a
measure at $\xi$ are that $N(F)$ actually be defined at $\xi$ for all $F\in C_c(\mathbb{G})$ and
that for each compact subset $K$ of $\mathbb{G}$ there is a constant $c_K$ so that for all $F\in C_c(\mathbb{G})$ with support inside $K$
we have $N(F)(\xi) \le c_K ||F||_\infty$. In fact this is precisely
the definition of measure via the Riesz representation theorem.
If for some $\xi\in X$, $N(F)(\xi)$ is defined, {\em positive}, and finite for each positive
$F\in C_c(\mathbb{G})$, then the right-hand side of \eqref{measureInterpretation2} provides a positive linear functional on $C_c(\mathbb{G})$ and so defines a positive measure.
\subsection{Some open questions}
There are numerous questions that remain to be investigated. Amongst these
we can point out:
\begin{itemize}
\item Are there simple process theoretical conditions on pure point measures $\omega$
that allow one to conclude that the processes atttached to $\omega$ can be
interpreted as measure valued processes?
\item Our homometry theorem shows that the stationary processes with fixed pure point diffraction form a group. Does this have any deeper meaning? Can we directly realize multiplication / inversion of stationary processes?
\item When there are moments of all orders, these moments characterize the corresponding
processes. There are various scenarios under which not all these moments are necessary to solve the inverse problem. For instance in \cite{XR} it is seen that real model
sets with real internal spaces are recoverable from their second and third moments. In \cite{LM}
it is seen that if $\mathcal{S} + \cdots +\mathcal{S} = \mathcal{E}$ ($m$ summands) then one needs only
the first $2m+1$ moments to solve the inverse problem. This suggests that solutions to the
inverse problem can be organized into hierarchies depending on the number of moments
required to specify them.
\item The sequence of moments suggests that there is some sort of cohomology theory
that would allow one to organize the increasing information that is added as successive
moments are included in the picture. Such a cohomology theory would be a most welcome
addition to this study.
\end{itemize}
|
\section{Introduction}
\label{intro}
Although several thousand gamma-ray bursts (GRBs) have been observed so far,
the mechanisms responsible for the GRB emission remain elusive. Both synchrotron and
photospheric models have been widely explored in the literature to explain the
characteristic ``Band-like'' GRB spectrum which is characterized by smoothly
connected power-laws (Band et al. 1993). The synchrotron models rely on a power-law
electron distribution accelerated at large distance (or small optical depths) in the jet.
The photospheric models focus, instead, on radiation that is released when the jet becomes transparent.
Photospheric models are attractive interpretation because they can result in high radiative efficiency
and naturally predict peak energies $E_{\rm peak}\sim1$ MeV close to
the observed values (Goodman 1986; Thompson 1994; M\'esz\'aros \& Rees 2000).
The value of $E_{\rm peak}$ in this interpretation is coupled to the main
properties of the flow (e.g. luminosity $L$, and Lorentz factor $\Gamma$).
Indeed, observations indicate that the peak of the $E f(E)$ spectrum
tracks the instantaneous gamma-ray luminosity and integrated energy
during a burst and among different bursts, respectively (e.g., Amati et al. 2002; Yonetoku
et al. 2004; Ghirlanda, Nava \& Ghisellini 2010).
On the other hand, the energy dissipated in the base of the jet effectively
thermalizes, so in the absence of additional dissipation at modest
optical depths (i.e. further out) the emission from the transparency radius is quasi-thermal
in sharp contrast to that typically observed. Significant dissipation of energy of some sort is
required close to the photosphere of the flow to lead to the observed,
smoothly-connected power-law spectra. The source of such dissipation
may be (strong or many, weak) shocks (Rees \& M\'esz\'aros 1994;
Lazzati \& Begelman 2011), magnetic reconnection (Giannios 2006),
or nuclear collisions (Beloborodov 2010).
Adopting the reconnection model for GRBs of Drenkhahn (2002),
I have shown that magnetic dissipation leads to powerful photospheric emission
(Giannios 2006). The observed $\sim$MeV peak of the spectrum forms at
Thomson optical depth of several tens where radiation and electrons drop out of thermal
equilibrium; the electrons turn hotter further out in the flow.
Inverse Compton scattering at larger distance (smaller optical depth)
leads to the high-energy tail that can extend well into
the GeV range. A Band-like spectrum naturally forms in this scenario.
Here, I develop a generic dissipative photospheric model applicable to
arbitrary dissipative process that results in electron heating.
Both a flat rate of dissipation of energy
(e.g. constant luminosity dissipated per decade of distance)
and localized dissipation events are explored (Sections 2, 3).
The model predicts a relation of the peak $E_{\rm peak}$ of the observed
emission with the properties of the flow;
most sensitively depending on the bulk Lorentz factor
$\Gamma$ (Section 4). Using the observed $E_{\rm peak}-L_{\gamma}$
relation I make inferences for the central engine of GRBs (Section 4.1).
Section 5 clarifies various aspects of the GRB variability in the context
of the model. Discussion and conclusions are presented in Section 6.
\section{Dissipative Photospheres}
\label{sec:disruption}
Energy dissipated at the very inner parts of the jet flow quickly thermalizes.
A substantial thermal component can also be built in an initially rather
cold, magnetically dominated flow when magnetic energy is dissipated
at large Thomson optical depth $\tau\gg 1$, e.g. as expected in a striped wind
model (Drenkhahn \& Spruit 2002). The thermal luminosity $L_{\rm th}$
(dominated by radiation) is released at the
transparency radius (defined as the distance at which $\tau=1$).
For typical parameters of the jet flow, the resulting quasi-thermal emission
peaks at $E_{\rm peak}\sim 3k_{\rm B}T_{\rm obs}\sim 0.1-1$ MeV
(Goodman 1986, Thompson 1994; Daigne \& Mochkovitch 2002;
Beloborodov 2010), very close to the typically observed values (Band
1993). In the absence of dissipation of energy close to the
photosphere, however, the emerging emission cannot account for
the observed GRB spectrum. Though isolated cases
for a strong quasi-thermal component in the GRB emission
have been made (Ryde 2005; Ryde et al. 2010; Ryde et al. 2011),
{\it the GRB spectrum generically has non-thermal appearance}.
The photospheric emission is, however, modified when additional
energy release takes place close to the transparency radius.
It turns out (see next Section) that continuous electron heating
at a range of optical depths from $\tau\sim$several tens out to
$\sim$0.1 may be the key to reproduce the observed emission. Continuous dissipation
results in a well defined distance where radiation and particles drop
out of equilibrium, the so-called {\it equilibrium radius}.
The peak of the emission spectrum is determined
by the plasma temperature at this distance.
Below we develop a general framework to calculate
the peak of the spectrum as function of the properties of the flow.
\subsection{A generic model}
\label{sec:jet}
Consider a jet coasting with bulk Lorentz factor $\Gamma$, total
isotropic equivalent luminosity $L$ and baryon
loading $\eta\equiv L/\dot{M}c^2\simmore \Gamma$. In the presence of energetic particles
injected by the dissipative process, the flow can be loaded with
a modest number of pairs
(Pe'er et al. 2006; Vurm et al. 2011). Assuming
$f_{\pm}$ electron+positron pairs per proton (i.e. $f_\pm =1$ for
$e-p$ plasma; hereafter pairs are referred to as electrons\footnote{
For a photospheric GRB model where the flow does not contain baryons
see Ioka et al. (2011).}), the rest-frame electron number density is
$n_{\rm e}=f_{\pm} L/4\pi r^2\eta\Gamma
m_{\rm p}c^3$. Clearly both $\Gamma$ and $f_{\pm}$ can vary with
distance (because of acceleration of the flow and energetic particle
injection that result in pair creation, respectively).
Here, I treat these quantities as constants with the main focus been
on their values close to the equilibrium radius.
The Thomson optical depth $\tau\equiv n_{\rm e}\sigma_{\rm T}r/\Gamma$
as function of distance is
\begin{equation}
\tau=37\frac{L_{53}f_{\pm}}{r_{11}\eta_{2.5}\Gamma_{2.5}^2},
\end{equation}
where all quantities are in cgs units and the $A=A_x10^{x}$ notation
is adopted. Setting $\tau=1$, the Thomson photosphere is located at
\begin{equation}
r_{\rm ph,11}=37\frac{L_{53}f_{\pm}}{\eta_{2.5}\Gamma_{2.5}^2}.
\end{equation}
We consider a flow which carries a thermal component
of luminosity $L_{\rm th}$ that is a substantial fraction $\epsilon$ of that of the flow:
$L_{\rm th}=(4/3)4\pi r^2\Gamma^2aT_{\rm th}^4c=\epsilon L$.
This can be realized in both fireballs and dissipative, magnetically
dominated flows (e.g. Drenkhahn 2002)
The (rest frame) plasma temperature is
\begin{equation}
k_{\rm B}T_{\rm th}=1.1\frac{\epsilon^{1/4}L_{53}^{1/4}}{r_{11}^{1/2}\Gamma_{2.5}^{1/2}}\quad \rm keV.
\end{equation}
Inspection of eqs. (2) and (3) reveals that
the plasma typically cools to sub-keV temperature close to the photosphere.
It turns out, however, that the dissipative process heats up the flow
to a temperature well in excess of $T_{\rm th}$ before it reaches the photosphere.
Suppose that a dissipative process injects energy in the flow heating the electrons
(see Section 3 for discussion on the physical justification of such assumption).
I consider two different situations for the radial dependence
of the dissipation rate: i) a gradual rate of energy release of rather
flat profile (i.e. constant rate of energy dissipated per e-folding of distance):
$dL_{\rm d}/dr=L_{\rm d,o}/r$ and ii) dissipation that is localized to a narrow range
in distance.
In the case of the gradual energy release, the heating rate per unit
volume in the rest frame of the flow is
$P_{\rm h}=L_{\rm d,o}/4\pi \Gamma r^3$. If the thermal component
$L_{\rm th}$ is built by the dissipation of energy from smaller distances
(and including adiabatic cooling of the photons)
one gets $L_{\rm th}=(3/2)L_{\rm d,o}$. Allowing for a comparable amount
of additional heating even deeper in the flow, we set $L_{\rm d,o}=L_{\rm th}/3=\epsilon
L/3$.
Heating of the electrons is balanced by radiative cooling\footnote{Adiabatic
cooling can be shown to be negligible for the electrons.}.
Electrons and photons are in thermal equilibrium
at large depth with their common temperature given by
eq.~(3)\footnote{In reality, the electrons maintain a slightly higher
temperature from radiation because of external heating.}.
This equilibrium is broken once heating and cooling balance
leads to $T_{\rm e}>T_{\rm th}$. It can be shown (see Giannios 2006) that Compton cooling
$P_{\rm IC}=4n_{\rm e}\Theta_{\rm e}c\sigma_{\rm T}U_{\rm r}$ is the dominant
cooling mechanism for the electrons in this region ($\Theta_{\rm
e}=k_{\rm B}T_{\rm e}/m_{\rm e}c^2$ and $U_{\rm r}$ is the energy
density of radiation).
Equating dissipative heating and cooling ($P_{\rm h}=P_{\rm IC}$)
and setting $U_{\rm r}=aT_{\rm th}^4$ one derives the electron temperature as
function of distance
\begin{equation}
k_{\rm B}T_{\rm e}=1.5\frac{r_{11}\Gamma_{2.5}^2\eta_{2.5}}{f_{\pm}L_{53}}\quad \rm keV.
\end{equation}
The location where radiation and electrons drop out of equilibrium is
found by setting $T_{\rm th}=T_{\rm e}$, defining the (maximum) {\it equilibrium radius} $r_{\rm eq}$:
\begin{equation}
r_{\rm eq,11}=0.80\frac{L_{53}^{5/6}\epsilon^{1/6} f_{\pm}^{2/3}}{\Gamma_{2.5}^{5/3}\eta_{2.5}^{2/3}}.
\end{equation}
At the equilibrium radius $r_{\rm eq}$, the Thomson optical depth and the temperature of the electrons are, respectively
\begin{equation}
\tau_{\rm eq}=46\frac{L_{53}^{1/6} f_{\pm}^{1/3}}{\epsilon^{1/6}\Gamma_{2.5}^{1/3}\eta_{2.5}^{1/3}}
\end{equation}
and
\begin{equation}
k_{\rm B}T_{\rm eq}=1.2\frac {\epsilon^{1/6}\Gamma_{2.5}^{1/3}\eta_{2.5}^{1/3}}{L_{53}^{1/6} f_{\pm}^{1/3}}\quad \rm keV.
\end{equation}
Note that the optical depth and temperature of the plasma at the equilibrium
radius depend very weakly on the flow parameters and are in the range of
several tens and $\sim 1$ keV, respectively.
The equilibrium radius is of particular importance since it is the location
at which the peak of the spectrum forms
under a wide range of conditions. The observed peak of the spectrum is
\begin{equation}
E_{\rm peak}\simeq \frac{4}{3}3\Gamma k_{\rm B}T_{\rm eq}\simeq 1.5
\frac {\epsilon^{1/6}\Gamma_{2.5}^{4/3}\eta_{2.5}^{1/3}}{L_{53}^{1/6} f_{\pm}^{1/3}}\quad \rm MeV.
\end{equation}
At distance $r>r_{\rm eq}$, and using eqs. (1) and (4),
the Compton $y$ parameter\footnote{For relativistically expanding
fluid the number of scatterings scales as $\sim \tau$ and not
$\propto \tau^2$ as in a static medium (see Giannios 2006).} is found
to be $y=4\Theta_{\rm e}\tau \sim 0.4$
{\it independently} of distance or the parameters of the flow leading
to significant up-scattering of the photons.
The Compton parameter is close to unity because the incoming
radiative luminosity $L_{\rm th}$ and dissipation rate $L_{\rm d,o}$ are comparable.
Up-scattering of the radiation emerging from $r_{\rm eq}$
at larger distance leads to broader spectra and to a flat high-energy tail
above $E_{\rm peak}$. The high-energy tail is followed by an exponential
cutoff at the energy that corresponds to the electron temperature
at the distance where the dissipation stops (see next section).
\subsection{Numerical Results}
The analytical arguments presented in the previous Section are fully supported by detailed Monte-Carlo
radiative transfer simulations. The code developed in Giannios (2006) is used to
study the electromagnetic spectrum emerging from a dissipative
photosphere. The inner boundary of the calculation is set at the
equilibrium radius where photons following a thermal distribution
of temperature $T_{\rm eq}$ are injected. The photons
are followed throughout the photosphere until the flow reaches an
optical depth of $\tau=0.1$ where the outer boundary is set.
The calculation includes Compton scattering, relativistic effects, while
the electron temperature is iterated until the heating
rate is matched by Compton cooling everywhere in the flow.
We do not include synchrotron emission and
non-thermal particle acceleration. While both are important in
determining the exact spectrum, they are model dependent.
For this paper, however, it is important that the peak of the
emission is set at the distance where radiation and
matter drop out of equilibrium and rather independently of the details
of non-thermal processes. In the following,
both extended dissipation with $dL_{\rm d}/dr=L_{\rm d,o}/r$
and localized dissipation are investigated.
\subsubsection{Extended Dissipation}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[angle=270]{figures/diss_photospheres.ps}}
\caption[] {Numerically calculated spectra at different optical depths in the jet
for the reference values of the parameters and for extended dissipation of energy.
The thermal injected photon spectrum at the equilibrium radius (black line) evolves into
a broader spectrum at the $\tau=1$ surface (red line) and develops a
flat power-law tail at the outer radius (corresponding to $\tau=0.1$; green line).
The peak of the $E f(E)$ emerging spectrum is determined at the equilibrium radius.}
\label{fig:diagram}
\end{figure}
Assuming extended dissipation of energy,
in Fig.~1 I plot the radiation spectrum at various distances of the flow
for the reference values of the parameters ($L=10^{53}$ erg$\cdot$s$^{-1}$, $\eta=\Gamma=300$,
$\epsilon=0.3$, $f_{\pm}=1$). Note that the thermal emission
at $r_{\rm eq}$ evolves into one of non-thermal appearance when passing through the $\tau=1$
surface building a flat high-energy tail. Inverse Compton scattering
with Compton $y$ parameter $y\simless 1$ results in a flat ($E f(E)\sim E^0$) emission above the peak.
The outer boundary of the calculation is set at larger distance
that corresponds to $\tau=0.1$. The high-energy cutoff $E_{\rm cut}$ is determined by
the temperature of the electrons at the outer boundary (in this example
$k_{\rm B}T_{\rm e, out}\simeq 300$ keV resulting in $E_{\rm cut}\simeq \Gamma k_BT_{\rm e, out}\simeq 100$
MeV). While in this paper, I set the outer boundary of dissipation by hand at $\tau=0.1$,
more detailed models, such as the magnetic reconnection model of
Drenkhahn \& Spruit (2002) predict where this cutoff takes place.
In Fig.~2 I show the resulting spectra for different values of the parameters.
The peak of the emission $E_{\rm peak}$ is very close to the analytic expression (8).
The overall shape of the spectra is very similar with only the peak and cutoff
energies depending on the parameters. A Band-like spectrum is
reproduced for very different parameters of the flow. The high-energy
tail has spectral slope $f(E)\propto E^{\beta}$ with $\beta\simless -1$
as observed. Below the peak energy the slope, as measured in the 10-100 keV
range, is $f(E)\propto E^{\alpha}$ with $\alpha \sim 0-1$. The slope
$\alpha$ is consistent but in the rather hard range in
comparison with the observed distribution. Note, however, that synchrotron and
synchrotron-self-Compton emission from larger distances (not included
in this work) can soften the spectrum below the peak (Giannios 2008;
Vurm, Beloborodov \& Poutanen 2011).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[angle=270]{figures/diss_phot_spectra.ps}}
\caption[] {Emerging spectra for different values of the parameters and for
extended dissipation of energy. The spectrum
has a Band-like shape practically independently of the luminosity, baryon loading
and Lorentz factor of the flow. The peak of the spectrum follows the scaling predicted by
eq. (8). The high energy cutoff is set by the location of the outer boundary of the simulation.}
\label{fig:diagram}
\end{figure}
\subsubsection{Localized Dissipation}
The process that dissipates energy in the jet may result in energy release over a narrow
range of distances. Here, I investigate how a thermal component carried
by the flow is modified depending on the optical depth at which such energy release takes place.
I assume that the thermal component $L_{\rm th}=\epsilon L$ is reprocessed by Compton
up-scattering through a narrow region of hot electrons. The electrons
are heated at a rate $L_{\rm d,o}=L_{\rm th}$ over a range of Thomson
depth $\tau_{\rm diss}...\tau_{\rm diss}/2$
(i.e. a factor 2 in distance). The electron temperature is determined
by heating-cooling balance. Setting the various parameters to their reference values,
Fig.~3 shows the emerging spectrum for different values of $\tau_{\rm diss}$.
Dissipation at large optical depths $\tau_{\rm diss}\simmore 10$
leads to a broadened, Planck-like distribution. The resulting
narrow emission spectrum has a steep rise/decline below/above the
peak. Dissipation at $\tau_{\rm diss}\simless 1$ leads to a hot
region above the photosphere. Most of the photons do not interact with the hot electrons
leading to a distinct quasi-thermal component while; photons
that are up-scattered at least once form a high-energy component.
Although occasionally GRB spectra show
such multi-component behavior in the keV-MeV regime, this in not typical.
For $\tau_{\rm diss}\sim 3$ the two components are smoothly connected and
the resulting spectrum compares more favorable with observations.
For a model where the dissipation (through magnetic reconnection)
takes place at a narrow region of $\tau\simmore 1$ see Thompson (1994).
Summarizing, localized dissipation in general does not account for
observations because it either leads to narrow or multi-component
spectra in the $\sim$MeV energy range.
There is a limited range of optical depths of $\tau_{\rm diss}\sim 3-5$
where dissipation results in a smooth Band-like spectrum.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[angle=270]{figures/tau_diss.ps}}
\caption[] {Emerging spectrum for dissipation of energy at a narrow range
of distance (or optical depth $\tau_{\rm diss}$) for different $\tau_{\rm diss}$.
Dissipation at large optical depths $\tau_{\rm diss}\simmore 10$
leads to narrow emission spectrum. Dissipation at $\tau_{\rm diss}\simless 1$ leads to distinct
thermal and high-energy components. For $\tau_{\rm diss}\sim 3$ the two components are
smoothly connected.}
\label{fig:diagram}
\end{figure}
\section{The dissipative mechanisms}
In the previous analysis we simplified the calculation
assuming thermal electrons continuously heated
by an external agent. Here we elaborate on the
possible sources for smooth ``volume" heating
of the GRB flow (Ghisellini \& Celotti 1999;
Stern \& Poutanen 2004; Pe'er, M\'esz\'aros
\& Rees 2006).
Models for gradual energy release that heats
the electrons involve magnetic reconnection (Giannios 2006),
multiple weak shocks (Ioka et al. 2007;
Lazzati \& Begelman 2010)\footnote{The, more often
invoked for the GRB emission, strong internal shocks
(Rees \& M\'esz\'aros 1994) are unlikely to lead to
smooth/gradual heating of the electrons. Such shocks, instead,
accelerate particles practically instantaneously to ultrarelativistic
energies leaving them to cool radiatively on a longer time scale
(see, however, Ramirez-Ruiz 2005).}
and neutron-proton collisions (Beloborodov 2010).
In the first case, the energy is initially stored in the
magnetic field while in the latter cases in relative
bulk motions within the jet. In the following I argue
that, independently of the dissipative mechanism,
a large fraction of the energy dissipated close to the
equilibrium radius is expected to be stored
to (mildly) relativistic protons. Coulomb $e-p$ collisions
effectively drive the energy from protons into the electrons
(see Beloborodov 2010).
The electrons then transfer their energy into the photon field
through Compton scattering. The characteristic
electron equilibrium temperature is sub-relativistic
and determined by heating-cooling balance.
Multiple weak shocks in the jet or elastic neutron-proton collisions
(close to the distance where the neutron and proton fluids decouple)
can be expected to heat the protons at mildly relativistic
temperatures with the heating maintained over a range
of distances in the jet. When the heating takes place at
optical depth of tens, the Coulomb coupling of
$\simless$ GeV protons with
$\sim$keV electrons (see, e.g., Stepney 1983) can be shown to be effective in transferring
the energy into the electrons within one expansion timescale of the jet.
These, slowly heated, electrons pass their energy into the photon field
effectively through inverse Compton scattering.
In addition to mildly relativistic protons, shocks and inelastic $n-p$
collisions can inject high-energy particles. Non-thermal
processes can result in a modest pair loading in the jet
(Pe'er et al. 2006; Beloborodov 2010)\footnote{Note, however,
that if the energy density of the magnetic field is higher than
that of the radiation, ultrarelativistic electrons
cool mainly through synchrotron emission and the pair creation is
further limited.}
but not expected to directly affect the peak energy of the emission as long they
are not energetically dominant.
Magnetic reconnection can take place effectively
at large optical depths (deep in the jet) through, e.g., tearing instabilities
of the current sheet (Loureiro, Schekochihin \&
Cowley 2007; Uzdensky, Loureiro \& Schekochihin 2010) assisted/induced
by the acceleration of the jet (Lyubarsky 2011).\footnote
{A possible speed up of the reconnection rate at Thomson thin
conditions (Lyutikov \& Blandford 2003; McKinney \& Uzdensky 2012)
can have interesting implications but is not
required in this picture.}
Magnetic reconnection takes place in multiple
dissipative centers that accelerate particles,
heat the plasma and drive fast bulk motions.
The fast motions in the downstream of the reconnection
region are required for sufficiently fast reconnection.
The bulk motions are dissipated further downstream
in shocks resulting in hot protons. Coulomb collisions
can effectively couple the proton energy to the electrons which,
in turn, couple it to the radiation field (as discussed above).
In this strongly magnetized plasma ($U_{\rm B}\simmore U_{\rm ph}$),
sub-relativistic electrons can thermalize very efficiently through exchange of synchrotron
photons (the so-called ``synchrotron boiler''; Ghisellini et al. 1998)
before they have the chance to cool through inverse Compton scattering.
An alternative picture where the energy released
in the reconnection process drives MHD waves
(instead of bulk motions) is described in Thompson (1994).
Photons extract the energy from the waves through
scattering on electrons (Compton drag).
Also in this picture the electron velocity is limited to
sub-relativistic speed and close to a Compton equilibrium
with the radiation field. The analysis of the previous
section may apply to this scenario as well.
In summary the details of the dissipation process
(and the related non-thermal processes) may well
affect the pair injection rate and result in additional features
in the spectrum (e.g. powering a pair
annihilation line at $E\sim \Gamma m_{\rm e} c^2$)
but do not change the basic picture of
sub-relativistic electrons strongly coupled to the
radiation field with their energy determined by a balance
of dissipative heating and radiative cooling. It is this basic
process at the equilibrium radius that determines the peak of the emerging
emission.
\section{Inferring the bulk Lorentz factor of the flow}
In this rather general framework of dissipative photospheres,
there is a close connection between the peak of the spectrum
and the properties of the flow summarized in eq.~(8). Eq.~(8)
can be solved in terms of $\Gamma$:
\begin{equation}
\Gamma=280 E_{\rm peak,MeV}^{3/5}\epsilon_{-0.5}^{-1/10}L_{53}^{1/10}
f_{\pm}^{1/5}\big(\frac{\eta}{\Gamma}\Big)^{-1/5},
\end{equation}
where I chose to express $\Gamma$ as function of the ratio $\eta/\Gamma\simmore 1$
instead of $\eta$. A clear implication from the last expression
is that {\it $\Gamma$ depends mainly on the
peak energy $E_{\rm peak}$ with extremely weak dependence on the
physical properties of the jet}. To first approximation,
one can have a fair estimate of the Lorentz factor
just from the observed peak of the GRB emission.
From eq.~(9) it is clear that the model predicts that
the high-peaked GRBs come from faster jets.
This is in sharp contrast to the synchrotron
internal shock model where the opposite holds
true ($E_{\rm peak}\propto \Gamma^{-2}$).
One can somewhat improve on the estimate
of the bulk $\Gamma$ of the jet from observables by assuming a
gamma-ray luminosity $L_{\gamma}\sim \epsilon L$
and using the observed $L_{\gamma}$ and $\epsilon\sim 1$.
This interpretation of the peak of the emission implies that
the brightest GRBs with $E_{\rm peak}\sim$ several MeV
and $L_{53}\sim 10$ (e.g. Abdo et al. 2009)
come from the most relativistic $\Gamma \sim 1000$
flows while weaker X-ray flashes with $E_{\rm peak}\sim 30$ keV
and $L_{53}\sim 0.01$ come from ``slower"
jets of $\Gamma\sim 20$. Low-luminosity GRBs
(e.g. Soderberg et al. 2006) and the X-ray flares
that follow many bursts (Burrows et al. 2005;
Chincarini et al. 2010) can also be a result of a
dissipative photosphere from yet slower jets of $\Gamma\simless 10$.
In particular low-luminosity GRBs may come
form $\Gamma\sim 3$ jets that result in only modest
relativistic beaming of their emission and may, therefore, account for the
larger observed local rate of these events
(Soderberg et al. 2006) in comparison to classical GRBs.
As discussed in Section 6, observational estimates of
the Lorentz factor of GRBs support the model prediction that higher
$E_{\rm peak}$ bursts are coming from higher $\Gamma$ flows.
\subsection{Additional inferences for the flow using observed
correlations of the GRB emission}
It has been recognized for some time that various {\it time
integrated} quantities over the duration of a burst may
correlate with each-other, e.g., the peak energy $E_{\rm peak}$ with
the isotropic gamma-ray energy $E_{\gamma}$ (Amati et al. 2002).
Even in the presence of outliers (Band \& Preece 2005; Nakar \& Piran
2005), these correlations
may teach us a lot about the GRB physics.
The emission in the photospheric models
is, however, connected to the {\it instantaneous} properties of
the flow. A change in any of the properties
(e.g. luminosity or baryon loading) during the burst shifts the location
of the photosphere and its appearance. Ghirlanda
et al. (2010; 2011) found that there is a time dependent
Amati-like relation of $E_{\rm peak}(t)$ and $L_{\gamma}(t)$ during
the evolution of bursts where $E_{\rm peak} \simeq 1L_{\gamma,53}^{1/2}$ MeV.
If this relation is verified by more data, it implies for the photospheric models that
the Lorentz factor correlates with the luminosity of the flow.
In this interpretation, one can derive an estimate of the Lorentz factor
of the flow directly from the observed gamma-ray luminosity
using eq. (9), the instantaneous $E_{\rm peak}-L_{\gamma}$ correlation
and that $L_{\gamma}\sim L_{\rm th}=\epsilon L$:
\begin{equation}
\Gamma=310 E_{\rm peak,MeV}^{4/5}\epsilon_{-0.5}^{-1/5}f_{\pm}^{1/5}\Big (\frac{\eta}{\Gamma}\Big)^{-1/5}.
\label{gammati}
\end{equation}
Note that the Lorentz factor depends sensitively on a single
observable, namely, the peak energy. Therefore, $E_{\rm peak}$
can be use the infer the Lorentz factor.
In terms of properties of the flow, I find that $\Gamma= 200 \epsilon_{-0.5}^{1/5}L_{53}^{2/5}
(\eta/\Gamma)^{-1/5}f_{\pm}^{1/5}$, i.e., that
more luminous bursts are less baryon loaded
with $\Gamma\propto L^{2/5}$.
Note, however, that a systematic dependence of any other quantity
with, say, the luminosity of the flow can distort this $\Gamma-L$ scaling.
Finally, the $\Gamma-L$ relation can be combined with eq.~(7) to find
the temperature of the flow at the equilibrium radius:
\begin{equation}
k_{\rm B}T_{\rm eq}\simeq 0.8 \epsilon_{-0.5}^{3/10}L_{53}^{1/10} f_{\pm}^{-1/5}\Big (\frac{\eta}{\Gamma}\Big)^{1/5}\rm \quad keV.
\end{equation}
Note that the temperature of the flow clusters in the $\sim$keV range
and has very weak dependence on the parameters of the jet. As
discussed in Section 6 such clustering of the peak of the
emission in the rest frame of the flow has been observed.
The instantaneous $E_{\rm peak}-L_{\gamma}$ relation also puts interesting constraints on the distance
where the peak of the spectrum forms. Using the expression (5) for
$r_{\rm eq}$ and eq.~(\ref{gammati}), I arrive to
\begin{equation}
r_{\rm eq}=2.1\times 10^{11} L_{53}^{-1/10}\epsilon_{-0.5}^{-3/10}f_{\pm}^{1/5}\Big (\frac{\eta}{\Gamma}\Big)^{-1/5}\quad \rm cm.
\end{equation}
The equilibrium distance $r_{\rm eq}$ depends very weakly on the various parameters. Even allowing
for orders of magnitude variations in the luminosity and (as expected) more modest changes
in other parameters from burst to burst (or during the evolution of a burst), $r_{\rm eq}$
varies at most by a factor of a few.
\section{GRB variability}
GRBs are variable on timescales as short as milliseconds.
Their lightcurves are characterized by multiple pulses
(e.g. Fenimore et al. 1995) that typically last for seconds and
show characteristic spectrum that often
evolves from hard to soft (e.g. Hakkila \& Preece 2011) and/or tracks the
instantaneous flux of the flow (e.g. Ghirlanda et al. 2011). Can the photospheric
model account for the temporal GRB behaviour?
The small radius of emission in the photospheric models
allows for fast variability down to $t_{\rm v}\sim r_{\rm
ph}/2\Gamma^2c \sim $sub-msec timescales (depending
on the parameters of the flow; see also Giannios \& Spruit 2007).
Allowing the continuous heating out to optical depth $\tau \sim
0.1$ ($r_{\rm out}= 10 r_{\rm ph}$) can lead to the
multi MeV high-energy tail delayed by $ r_{\rm
out}/2\Gamma^2c\sim$several msec with respect to
the MeV peak (with the GeV emission potentially more delayed depending
on the distance at which it takes place). The steady-state jet model developed here
is not applicable to study extremely short (sub-msec) timescales
(for which the steady-state assumption breaks down) but is well suited to study
the evolution of the burst on longer timescales.
In particular, the $\sim$sec duration GRB pulses can be accurately studied
as a sequence of steady state models where the properties of the jet
($L=L(t)$, $\Gamma=\Gamma(t)$, etc)
vary on this timescale. The observed variability on second
timescales in the well studied $\sim 10-1000$ keV energy range
is likely to be dominated by temporal evolution of the properties
of the jet rather than delays introduced by propagation effects of the
ejecta.
The spectral evolution during a GRB pulse depends on how
$\Gamma(t)$ and $L(t)$ evolve during the pulse.
Giannios \& Spruit (2007) have shown that {\it If}, for
instance, $\Gamma(t)$ has a positive power-law dependence on $L(t)$
during individual pulses then the peak of the spectrum tracks the observed flux (as seen
observationally in Ghirlanda et al. 2011). In this case, while the
flux declines after the peak luminosity of a pulse, the peak of the emission
spectrum also declines and the spectrum softens. As a result the
pulse duration on the softer X-ray bands is longer than in the harder ones.
In summary, continuous dissipation close to the photosphere can
allow for fast evolving (sub-msec) emission. Furthermore, it is possible to account
for observed temporal properties of GRBs given specific assumptions
about the behaviour of the central engine. The reasons for which the
engine operates in such a fashion is, however, not addressed by this work.
\section{Discussion and Conclusions}
\label{sec:discussion}
In this paper, I have explored a wide range of
``dissipative photosphere'' models to pin
down the location where the peak $E_{\rm peak}$
of the spectrum forms and how it connects to the properties
of the jet. The resulting expression (9) of my analysis
summarizes this connection, is quite general and
rather independent of the physical model for energy
dissipation.
I have explored both localized and continuous in distance dissipation
of energy. Localized heating of electrons over a narrow
range of distance is shown here to have difficulties to
account for observations. If the dissipation takes place
at large optical depth, the emerging spectrum is rather
narrow, while dissipation at $\tau\simless 1$ has two
distinct components in the $\sim$MeV energy range, both
in conflict to the majority of observed bursts. However,
localized dissipation that takes place
at optical depth $\tau\sim 3$ results is a more promising
spectrum. Thompson (1994) discusses a scenario where
such dissipation is possible.
I find that {\it continuous} dissipation
of energy over a wide range of optical depths
can naturally give the Band-like spectrum with
peak and slopes above and below the peak at the observed range.
Magnetic reconnection in a jet that contains
small-scale field reversals naturally results in such flat dissipation
profile throughout the photospheric region. Alternatives
such as multiple, weak shocks or neutron-proton collisions
can also result in continuous energy injection.
The MeV peak of the spectrum forms at
the distance where radiation and the electrons
drop out of thermal equilibrium. In the context of this model
the peak energy is mainly determined by the
bulk Lorentz factor of the flow. The observed $E_{\rm peak}$
can, therefore, be used to infer $\Gamma$.
The model predicts that the peak energy $E_{\rm peak}$ positively correlates
with $\Gamma$. I find that weak
GRBs (the spectrum of which peaks in the X-ray regime; the so called
X-ray flashes) and X-ray flares that follow the bursts
potentially come from $\Gamma\sim 10$ jets while the
brightest observed {\it Fermi-LAT} bursts, with peak energy at several
MeV, come from the fastest $\Gamma\sim1000$ flows.
Recently (Ghirlanda et al. 2012) used the peak time of the afterglow
emission to estimate the bulk Lorentz factor $\Gamma$ of 31 GRBs.
They verified that brighter bursts are characterized by higher
$\Gamma$ and showed that the peak energy of the emission at the rest frame
of the jet clusters at several keV; both findings are unique predictions of
this work (see eqs. 7, 11).
The observed relation of the peak of the spectrum with the {\it
instantaneous} burst luminosity
indicates a close coupling of the emerging spectrum
to instantaneous properties of the flow (Ghirlanda et al. 2011).
It has already been pointed out in Giannios \& Spruit (2007) that
the $E_{\rm peak}-L_{\gamma}$ relation implies,
in the context of the reconnection model, that
brighter segments of a burst come from higher $\Gamma$
(i.e., ``the brighter the cleaner'') jet.
I come to the same conclusion for the generic photospheric
model developed here where the bulk Lorentz factor $\Gamma$
and the luminosity of the flow $L$ scale as, roughly, $\Gamma\propto L^{2/5}$.
Note again that independent constraints coming from afterglow modeling
indicate a similar correlation for $\Gamma$
and the gamma-ray energy $E_{\gamma}$ (Liang et al. 2010).
In this model, the MeV peak forms in a
fairly compact region, while the high-energy tail
forms at a larger
distance. The tail can extend to multi-GeV energy without
suffering attenuation due to pair creation and can exhibit delays of
the order of $r_{\rm GeV}/\Gamma^2c$ with respect to
the MeV emission that may
be of order of seconds (as observed in some {\it Fermi-LAT}
bursts, see Abdo et al. 2009; $r_{\rm GeV}$
is the radius where GeV emission takes place).
While a different mechanism may be invoked
for the dissipation at large distance that leads to the GeV emission
(M\'esz\'aros \& Rees 2011), the continuous spectrum from MeV to GeV
energy indicates a single dissipative mechanism\footnote
{but not necessarily a single emitting region
for the MeV and GeV emission.} (see, e.g., Bosnjak \& Kumar
2012).
Interestingly, using the observed instantaneous $E_{\rm
peak}-L_{\gamma}$ relation, the equilibrium distance $r_{\rm eq}$
(where the peak of the emission is set) can be shown to depend extremely weakly
on the parameters of the flow with $r_{\rm eq}\simeq 2\times 10^{11}$ cm.
This value is similar to the radii of Wolf-Rayet stars,
probable progenitors of long-duration GRBs. This could indicate that
interactions of the jet with the progenitor are important for
energy dissipation (e.g. Lazzati \& Begelman 2010).
On the other hand, a similar correlation is observed in
short-duration GRBs, supporting the idea
of a progenitor-independent dissipative mechanism
(Ghirlanda et al. 2011).
The inferred ``the brighter the cleaner" property of GRB
flows has profound implications for the nature of the central engine.
Both accreting black holes and rapidly rotating magnetars
have been invoked for launching the jet.
Models that invoke accretion into a black hole (Woosley 1993)
are not sufficiently developed to predict the amount of baryons
that make it into the jet. In the millisecond magnetar model
for GRBs (Usov 1992; Metzger et al. 2011), the calculation of the
baryon loading is more tractable. The magnetars born with the
strongest fields drive the
brightest bursts and also give, averaged over the GRB
duration, less baryon loaded flows. In particular, oblique rotators result in
$\Gamma \propto L^{0.6}$ (but with a large scatter; see
Metzger et al. 2011). This is a rather intriguing
result since it might point to a rather complete picture for GRB
physics (which is, however, hardly unique).
A central engine consists of a protomagnetar
(with its magnetic field axis, in general, misaligned to the rotational axis),
that gives rise to a magnetically dominated jet that contains
field reversals on small scale (striped wind).
Magnetic reconnection proceeds in the jet at a wide range of
distances from the central engine and at a rather flat rate (Drenkhahn
\& Spruit 2002). The bright MeV emission of the GRB emerges from the Thomson
photosphere of the flow while residual dissipation can extend
the high-energy tail above the MeV peak well into the GeV regime (Giannios 2008).
\section*{Acknowledgments}
DG acknowledges support from the Lyman Spitzer, Jr. Fellowship awarded
by the Department of Astrophysical Sciences at Princeton University
and from the Fermi 4 Cycle grant number 041305.
|
\section{Introduction}
Dust grains play crucial roles in many aspects of the interstellar medium (ISM).
For example, alignment of dust grains with respect to magnetic field provides
insight into star formation through far-infrared and submm polarized emission
(see Lazarian 2007 for a review).
Very small spinning dust grains radiate microwave emission that contaminate
to cosmic microwave background (CMB) radiation (Draine \& Lazarian 1998;
Hoang, Draine \& Lazarian 2010; Hoang, Lazarian \& Draine 2011).
Optical extinction and polarization properties of dust depend mainly on its
size distribution. Grain-grain collisions, which depend on grain relative
motions, govern grain coagulation and destruction, that result in the grain
size distribution (see Hirashita \& Yan 2009). Grain-grain collisions are
considered the first stage of planetesimal formation in circumstellar disks
(see e.g., Dullemond \& Dominik 2005).
Traditionally it is believed that the motion of dust grains in the ISM arises
from radiative force, ambipolar diffusion and hydrodrag (see Draine 2011).
The resulting motion from these processes is sub-Alfv\'{e}nic (i.e., $v\ll
V_{\A}=B/\sqrt{4\pi\rho}$ where $B$ is the magnetic field strength and
$\rho$ is the gas mass density), except in special environment conditions
(see Purcell 1969; Roberge et al. 1993).
Astrophysical environments are practically all magnetized and turbulent,
and turbulence is expected to be an important factor in accelerating dust
grains. The evidence for turbulence from electron density fluctuations
testifies for the existence of the so-called Big Power Law in the Sky
(Armstrong et al. 1995; Chepurnov \& Lazarian 2010), while the fluctuations
of velocity (see Lazarian 2009 and references therein) provide convincing
evidences of the dynamic nature of the observed inhomogeneities.
Whether an environment is thermally dominated or magnetized dominated
depends on the plasma $\beta$ parameter, which is defined as the ratio
of gas pressure to magnetic pressure
\begin{eqnarray}
\beta=\frac{8\pi n_{{\rm H}}k_{\rm B}T_{{\rm gas}}}{B^{2}}=0.1\left(\frac{n_{{\rm H}}}{30~\cm^{-3}}\right)
\left(\frac{T_{{\rm gas}}}{100~\K}\right)\left(\frac{10~\mu G}{B}\right)^{2},~~~
\end{eqnarray}
where $n_{{\rm H}}$ is the gas density and $T_{{\rm gas}}$ is the gas temperature.
Recent decade has been marked by substantial progress in understanding
of MHD turbulence. This included generalizing incompressible Alfv\'{e}nic
turbulence\footnote{While there are still ongoing debates about the detailed
structure and dynamics of incompressible MHD turbulence, we believe that
Goldreich \& Sridhar (1995) model provides an adequate starting point. In fact,
recent studies in Beresnyak \& Lazarian (2010) and Beresnyak (2011) provided
additional supports for the model.} by Goldreich \& Sridhar (1995) to realistically
compressible media and successful testing of the compressible theory (Lithwick
\& Goldreich 2001; Cho \& Lazarian 2002, 2003; Kowal \& Lazarian 2010). In
what follows in describing compressible MHD turbulence we shall be guided
by the mode decomposition of MHD turbulence into Alfv\'{e}n, slow and, fast
modes presented in Cho \& Lazarian (2002, 2003).
Studies of grain acceleration for magnetized turbulent environments were
initiated by Lazarian \& Yan (2002) who dealt with the acceleration by
incompressible Alfv\'{e}nic turbulence. Comprehensive studies of the acceleration
in realistically compressible environments were performed in Yan \& Lazarian
(2003, hereafter YL03) and Yan, Lazarian \& Draine (2004, hereafter YLD04).
Those studies identified gyroresonant interactions of grains with fast MHD
modes as a new powerful mechanism of grain acceleration. Recently,
Yan (2009) considered betatron acceleration and came to the conclusion
that for most environments the betatron acceleration is subdominant to
the gyroresonance acceleration for sub-Alfv\'{e}nic grains. The application
of grain velocities predicted by gyroresonance for modeling dust extinction
curve provided good correspondences between observations and theoretical
predictions (Hirashita \& Yan 2009) indicating that the turbulence is indeed
the main driving force behind grain acceleration.
The gyroresonance acceleration due to compressible MHD turbulence
(YL03;YLD04) was studied using quasi-linear theory (QLT, Jokipii 1966;
Schlickeiser \& Miller 1998). The underlying assumption of the QLT is that
the guiding center is assumed to move in a regular trajectory along a
uniform magnetic field $\Bv_{0}$. The condition for a grain with velocity $v$
to resonantly interact with fast MHD modes at the scale $k_{\|}$ is given by
$\omega-k_{\|}v\mu-\omega=n\Omega$, for $n=0, \pm 1,\pm2, ...$ where
$\mu$ is the cosine of the grain pitch angle between $\vv$ and $\Bv_{0}$,
$\omega$ is the wave frequency, and $\Omega$ is the Larmor frequency
of the charged grain around $\Bv_{0}$. The gyroresonance acceleration
with $n\ne 0$ is dominant by eddies with size equal to gyro radius $l\sim r_{g}$.
Transit-time damping (TTD) or transit-time acceleration, arises from resonant
interactions of particles with the compressive component of magnetic fluctuations
(i.e., the component parallel to the mean magnetic field $\Bv_{0}$).
When the grain moves together with the wave along $\Bv_{0}$, it is
subject to magnetic mirror forces $-(mv_{\perp}^{2}/2B)\nabla_{\|}\Bv$, where
$m$ is the grain mass, $v_{\perp}$ is the grain velocity component perpendicular
to $\Bv_{0}$, and $\Bv$ is the total magnetic field.
In the plasma reference, the back and forth collisions of the grain with
the moving magnetic mirrors increase grain energy because the head-on
collisions are more frequent than trailing collisions due to the larger relative
velocity between grain and wave (see Fisk 1976; Schlickeiser \& Miller 1998).
The TTD acceleration with resonance condition
$k_{\|}v\mu=\omega$, was disregarded in previous studies on grain acceleration
because grains are expected to move slowly along the uniform magnetic field,
for which they can not catch up with the propagation of magnetic mirrors
along this direction. Although the gyroresonance is found to be able to
accelerate large grains to super-Alfv\'{e}nic velocities (see YL03; YLD04),
which is sufficient to trigger TTD, the resulting grain motion mostly
perpendicular to the uniform field $\Bv_{0}$ (i.e. $v\mu=0$) in the
QLT regime makes TTD incapable.
Due to magnetic fluctuations in the ISM, the local magnetic field $\Bv$ can
be decomposed into a uniform field plus a turbulent component, i.e.,
$\Bv=\Bv_{0}+\delta \Bv$. Thus, any perturbation $\delta \Bv$ will induce
the fluctuations of grain guiding center from a regular trajectory along
the uniform magnetic field (see e.g., Shalchi 2005). Non-linear theory
(hereafter NLT) for gyroresonance that takes into account such fluctuations
of guiding center was formulated in Yan \& Lazarian (2008, hereafter YL08)
to describe the propagation of energetic particles, and later it was applied
in Yan et al. (2008, hereafter YLP08) to study acceleration of energetic
particles in solar flares.
The important modification present in the NLT is the broadening of resonance
function from a Delta function $\delta(\omega-k_{\|}v_{\|}-n\Omega)$ to a
Gaussian function $R_{n}(\omega-k_{\|}v_{\|}-n\Omega)$ (see YL08). Such a
broadening of the resonance condition allows grains moving with $v\sim V_{\A}$
perpendicular to $\Bv_{0}$ to have TTD with compressive waves propagating
along $\Bv_{0}$. We are going to clarify the effects that TTD induces on grain
acceleration in the present paper.
In what follows, we revisit the basics of resonance acceleration for charged
grains by taking into account additional physical processes that were not
considered within original treatments. In discussing the gyroresonance
acceleration, we are going to take into account the fluctuations of grain
guiding center. In particular, we are going to investigate the efficiency of
TTD on grain acceleration in MHD turbulence.
The structure of the paper is as follows. In \S 2, we present briefly the
problem of grain charging, important dynamical timescales, and identify
the range of grain size in which grain charge fluctuations are important.
We revisit gyroresonance acceleration, and introduce TTD acceleration
in \S 3. Grain velocities induced by gyroresonance acceleration and TTD
are presented in \S 4. \S 5 is devoted for stochastic acceleration by low
frequency Alfv\'{e}n waves. Discussion and summary are presented in
\S 6 and 7, respectively.
\section{Grain Charging and Dynamics}
\subsection{Grain Charging}
Charging processes for a dust grain in the ISM consist of its sticking collisions
with charged particles in plasma (Draine \& Sutin 1985) and photo-emission induced
by $h\nu\ge 13.6$ eV photons (Weingartner \& Draine 2001).
In the former case, the grain acquires charge by capturing electrons and ions
from the plasma, while in the latter case the grain looses charge by emitting
photoelectrons. After a sufficient time, these processes result in a statistical
equilibrium of ionization, and the grain has a mean charge, denoted by
$\langle Z\rangle$, which is equal to the charge averaged over time. Due to
the discrete nature of charging events, the grain charge fluctuates around
$\langle Z\rangle$. The probability of finding the grain with charge $Ze$ is
described by charge distribution function $f_{Z}$. Here we find the charge
distribution $f_{Z}$ using statistical ionization equilibrium as in
Draine \& Sutin (1985) and Weingartner \& Draine (2001). Hoang \& Lazarian (2011)
found that the statistical ionization equilibrium is not applicable for tiny grains
with size $a<10\Angstrom$ for which the charging is infrequent, but it
is adequate for grains considered in the present paper.
Figure \ref{timescale} shows the variation of the grain mean charge
$|\langle Z\rangle|$ for graphite and silicate grains in the cold neutral
medium (CNM), warm neutral medium (WNM), warm ionized medium
(WIM). In the WIM, $\langle Z\rangle$ varies rapidly with the grain size,
and change its sign at $a\sim 10^{-6}$ and $\sim 10^{-5}$ cm, marked by
filled circles.
Let us define a characteristic relaxation time of the charge fluctuations,
$\tau_{Z}$, which is equal to the time required for the grain charge to
relax from $Z$ to the equilibrium state (Draine \& Lazarian 1998b):
\begin{eqnarray}
\tau_{Z}=\frac{\langle (Z-\langle Z\rangle)^{2}\rangle}{\sum_{Z}f_{Z}J_{tot}(Z)}
\equiv\frac{\sigma_{Z}^{2}}{\sum_{Z}f_{Z}J_{\rm tot}(Z)},
\end{eqnarray}
where $J_{\rm tot}(Z)$ is the total charging rate due to collisional charging
and photoemission (see Draine \& Sutin 1987; Weingartner \& Draine 2001).
Here we averaged over all possible charge states $Z$ to find $\tau_{Z}$.
In an ambient magnetic field $\Bv$, the grain with mean charge $\langle Z\rangle e$
gyrates about $\Bv$ on a timescale equal to the Larmor period:
\begin{eqnarray}
\tau_{L}=\frac{2\pi m c}{|\langle{Z}\rangle| eB}=1.56\times 10^{2}\left(\frac{a}{10^{-6}~\cm}\right)\left(\frac{\mu G}{|\langle Z\rangle|B}\right) {\rm yr},~~~ \label{tauL}
\end{eqnarray}
where $m=4/3\pi a^{3}\rho_{d}$ with $\rho_{d}$ being the dust mass density is
the grain mass. We adopt $\rho_{d}=2.2$ and $3.0$ g $\cm^{-3}$ for graphite
and silicate grains, respectively. The Larmor frequency reads
$\Omega=(\langle{Z}\rangle eB)/mc$.
We calculate the relaxation time of charge fluctuations $\tau_{Z}$ for both
graphite and silicate grains in various phases of the ISM with physical
parameters listed in Table 1. Figure \ref{timescale} compares $\tau_{Z}$
with the gas drag time $\tau_{{\rm drag}}$ (see Eq. \ref{tau_drag}) and the
Larmor period $\tau_{L}$. It can be seen that $\tau_{Z}\ll \tau_{L}<\tau_{{\rm drag}}$
for grains larger than $\sim 2\times 10^{-7}$ cm. For grains smaller than
$\sim2\times 10^{-7}$ cm, $\tau_{Z} \ge \tau_{L}$, so that the assumption
for grains to have a constant charge is no longer valid. As a result, the
fluctuations of grain charge should be accounted for in the treatment of
resonance acceleration for such very small grains. This issue will be
addressed in our future paper, in which we employ Monte Carlo method to
simulate grain charge fluctuations (see e.g., Hoang \& Lazarian 2011). In the
present paper, for the sake of simplicity, we adopt $\langle Z\rangle e$
for grain charge within the entire range of the grain size distribution.
\begin{figure}
\includegraphics[width=0.45\textwidth]{f1.ps}
\caption{Mean grain charge $|\langle Z\rangle|$ as functions of grain size
$a$ for graphite and silicate grains in different ISM phases. For the WIM,
$|\langle Z\rangle|$ changes rapidly with $a$, and filled circles mark
the change in grain charge between being positively charged (+) and
negatively charged (-). }
\label{sigmaZ}
\end{figure}
\begin{figure}
\includegraphics[width=0.45\textwidth]{f2.ps}
\caption{Timescales for gas drag $\tau_{{\rm drag}}$, Larmor period $\tau_{L}$,
and charge fluctuations $\tau_{Z}$ as functions of the grain size $a$ for
subsonic silicate grains in the various phases of the ISM. Shaded area
marks the range of grain size in which the charge fluctuations are important,
i.e., $\tau_{Z}\geq \tau_{L}$. The peaks in $\tau_{L}$ correspond to
the change in sign of grain charge.}
\label{timescale}
\end{figure}
\subsection{Grain Translational Damping}
Interactions of dust grains with the ambient gas present the primary mechanism
of dissipating translational motions of grains. The damping rate of translational
motion arising from the interaction with neutral gas is essentially the inverse
time for collisions with the mass of the gas equal that of a grain (Purcell 1969),
\begin{eqnarray}
\tau_{dn}^{-1}&=&\sqrt{\frac8{\pi}}\frac{n_n}{a\rho_d}(m_{n}k_{\rm B}T_{n})^{1/2},\nonumber\\
&=&2.4\times10^{-12}\left(\frac{10^{-6}\cm}{a}\right)\left(\frac{n_{n}}{30\cm^{-3}}\right)
\left(\frac{T_n}{100~\K}\right)^{1/2}\s^{-1},~~~~
\label{tau_fr}
\end{eqnarray}
where $m_n$, $n_n$, and $T_n$ are the mass, volume density, and temperature
of neutrals, and $a$ is the grain radius.
When the ionization degree is sufficiently high, the interaction of charged
grains with the plasma becomes important. The ion-grain cross section due
to long-range Coulomb forces is larger than the atom-grain cross section. As
a result, the rate of translational motion damping gets modified. For subsonic
motions the effective damping time due to gas drag is renormalized:
\begin{eqnarray}
\tau_{{\rm drag}}=\alpha^{-1}\tau_{dn} \label{tau_drag}
\end{eqnarray}
with the following renormalizing factor (Draine \& Salpeter 1979)
\begin{eqnarray}
\alpha=1+\frac{n_{\rm H}}{2n_n}\sum_{i}x_i\left(\frac{m_{i}}{m_n}\right)^{1/2}
\sum_{Z}f_{Z}\left(\frac{Ze^{2}}{ak_{\rm B}T_i}
\right)^{2}\nonumber\\
\times\ln\left[\frac{3}{2|Z|\sqrt{\pi xn_{{\rm H}}}}\left(\frac{k_{\rm B}T_i}{e^2}\right)^{3/2}\right]
.\hspace{.5cm}\label{alpha1}
\end{eqnarray}
Here $x_{i}$ is the abundance of ion $i$ (relative to hydrogen) with mass
$m_{i}$ and temperature $T_i$, $x=\sum_i x_i$, $Ze$ is the grain charge, and
$f_{Z}(Z)$ is the grain charge distribution function. When the grain velocity $v_d$
relative to gas becomes supersonic, the dust-plasma interaction is
diminished, and the damping rate in this case is renormalized due to the
gas-dynamic correction (Purcell 1969),
\begin{equation}
\alpha=\left(1+\frac{9\pi}{128}\frac{v_d^2}{C_{\rm s}^2}\right)^{1/2},\label{alpha2}
\end{equation}
where $C_{\rm s}=\sqrt{k_{\rm B}T_{n}/m_n}$ is the sound speed.
If $\tau_{\rm L}$ is greater than $\tau_{{\rm drag}}$, then the effect of magnetic
field on dust dynamics is negligible. However, $\tau_{\rm L} \ll \tau_{{\rm drag}}$
in most phases of the ISM.
\begin{table}
\caption{Idealized Environments and MHD turbulence parameters}\label{ISM}
\begin{tabular}{llll} \hline\hline\\
\multicolumn{1}{c}{\it Parameters} & \multicolumn{1}{c}{CNM}&
{WNM} &WIM\\[1mm]
\hline\\
$n_{\rm H}$~(cm$^{-3}$) &30 &0.4 &0.1 \\[1mm]
$T_{\rm gas}$~(K)& 100 & 6000 &8000\\[1mm]
$x_{\rm H}$ &0.0012 &0.1 &0.99 \\[1mm]
$B~(\mu G)$ &6 &5.8 &3.35 \\[1mm]
$L~(\pc)$ &0.64 &100 &100 \\[1mm]
$\delta V=V_{\A}~({\rm km}~\s^{-1})$&{2} &20 &20 \\[1mm]
$k_{c}~({\cm}^{-1})$&{$7\times 10^{-15}$} &$4\times 10^{-17}$ &... \\[1mm]
{Damping}&{Neutral-ion} &{Neutral-ion} &{Ion~ viscous}\\[1mm]
& & & {and~ collisionless} \\[1mm]
\hline
\footnotetext{Here $n_{{\rm H}}$ is the gas density, $T_{{\rm gas}}$ is the gas temperature,
$x_{{\rm H}}$ is the ionization fraction of H, $B$ is the strength of magnetic field,
$L$ is the injection scale of turbulence, $\delta V$ is the rms velocity of turbulence
at the injection scale, and $k_{c}$ is the cutoff scale of turbulence due to
collisional and collisionless damping.}
\end{tabular}
\end{table}
\section{Resonance Acceleration}
\subsection{Gyroresonance acceleration: nonlinear theory}
In this section, we revisit the treatment of resonance acceleration by
fast modes in compressible MHD turbulence using nonlinear theory (NLT).
Consider a grain of mass $m$, charge $Ze$, moving with velocity $v$ in a
magnetized turbulent medium with a uniform magnetic field $\Bv=\Bv_{0}$.
The motion of such charged grain in $\Bv$ consists of the gyration of the
grain about its guiding center and the translation of the guiding center
along $\Bv$. In the QLT limit, the guiding center is assumed to follow a
regular trajectory along $\Bv$ with a constant cosine of pitch angle $\mu=\cos\beta$
with $\beta$ being the angle between $\vv$ and $\Bv$. Gyroresonant
interactions between grain and wave occur when the wave
frequency in a reference system fixed to the grain guiding center is
a multiple of the Larmor frequency:
\begin{eqnarray}
\omega-k_{\|}v\mu=n\Omega \label{eq:gyro_res}
\end{eqnarray}
with $n=\pm 1, \pm2, ...$. This resonance condition is equivalently described by
a Delta function $\delta_{n}(\omega-k_{\|}v\mu-n\Omega)$.
Gyroresonance accelerates grains in the direction perpendicular
to the mean magnetic field $\Bv_{0}$ because electric field induced by plasma
perturbations is perpendicular to $\Bv_{0}$ (see e.g. YLD04). This acceleration
mechanism is dominant by eddies smaller than the grain gyroradius,
i.e., $l\le r_{g}$. Indeed, consider gyroresonance by fast MHD modes in
low-$\beta$ plasma. From the resonance condition (\ref{eq:gyro_res}) for
$n=1$ with $\omega=k_{\|}v_{\A}$, we obtain turbulent scales for gyroresonance
$k\ge k_{\res}=r_{g}^{-1}$ or $l\le r_{g}$, where the fact that $\mu\ge -1$ has been used.
In a turbulent medium, the local magnetic field $\Bv=\Bv_{0}+\delta \Bv$
where $\delta \Bv$ is the turbulent component of magnetic field, varies both
in space and time, so $\mu$ changes, and $v_{\|}$ and $v_{\perp}$ change
accordingly. The grain guiding center has fluctuations from its regular trajectory
along $\Bv_{0}$. The NLT takes into account such fluctuations of the guiding
center.
Assuming that the projection of the fluctuations of grain guiding center onto
the mean field $\Bv_{0}$ can be described by a Gaussian distribution,
the resonance condition becomes
\begin{eqnarray}
R_{n}\left(\omega-k_{\|}v\mu-n\Omega\right)=\frac{\sqrt{\pi}}{k_{\|}\Delta
v_{\|}}{\rm exp}\left[-\frac{(k_{\|}v\mu-\omega+ n\Omega)^{2}}{k_{\|}^{2}
(\Delta v_{\|})^{2}}\right],~~~
\end{eqnarray}
where $n=0$ and $\pm 1$, and $\Delta v_{\|}$ is the dispersion of
velocity (see YL08 and YLP08; Appendix C).
We are interested in the grain acceleration, so the diffusion coefficient
arising from gyro-phase averaging $D_{pp}$ is used (see Schlickeiser
\& Miller 1998). In compressible MHD turbulence, the fast modes are shown
to be dominant in gyroresonance acceleration (YL03). Its corresponding
diffusion coefficient is given by (see Appendix C)
\begin{eqnarray}
D_{pp}(\mu,p)^{\rm G}=\frac{v\sqrt{\pi}\Omega^{2}(1-\mu^{2})m^{2}V_{\A}^{2}M_{\A}^{2}}
{4LR^{2}}\int_{1}^{k_{c}L}x^{-5/2}dx\nonumber\\
\times\int_{0}^{1}\frac{d\eta}{\eta\Delta \mu}[J_{0}^{2}(w)+J_{2}^{2}(w)]
{\rm exp}\left[-\frac{(\mu-\frac{V_{A}}{\eta v}+n\frac{1}{\eta xR})^{2}}
{(\Delta \mu)^{2}}\right],~~~~\label{eq:dpg}
\end{eqnarray}
where $n=\pm 1$. In the above equation, $L$ is the injection scale of
turbulence, $w={k_{\perp} v_{\perp}}/{\Omega}, x=k/k_{\min}=kL,
R=vk_{\min}/\Omega, M_{\A}^{2}=
\delta V^{2}/V_{\A}^{2}$, $\eta=\cos\theta$ with $\theta$ is the angle between
the wave vector ${\bf k}$ and the mean magnetic field, $k_{c}$ is the cut-off
of turbulence cascade due to damping, and $J_{n}$ is second order
Bessel function. The dispersion of the cosine of pitch angle $\Delta\mu$ is
given by Equation (\ref{eq:dmu}) in Appendix C.
\subsection{Transit-Time Damping (TTD)}
Transit-time damping (TTD) or transit-time acceleration, arises from
resonant interactions of particles with the compressive component of
magnetic fluctuations, i.e., the component parallel to the mean
magnetic field $\Bv_{0}$ in magnetized turbulent environments.
When the grain moves together with the wave along $\Bv_{0}$,
it is subject to magnetic mirror forces $-(mv_{\perp}^{2}/2B)\nabla_{\|}\Bv$,
where $v_{\perp}$ is the grain velocity component perpendicular to $\Bv_{0}$.
In the plasma reference, the back and forth collisions of the grain with
the moving magnetic mirrors increase grain energy because the
head-on collisions are more frequent than trailing collisions
(see e.g., Fisk 1976). The resonance condition for a
grain with velocity $\vv$ reads
\begin{eqnarray}
\omega-k_{\|}v_{\|}=0,\label{eq:resTTD}
\end{eqnarray}
where $v_{\|}=v\mu$ is the grain velocity component parallel to $\Bv$,
and $\omega$ is the wave frequency (see Fisk 1976; Schlickeiser \& Miller 1998).
For fast MHD modes in low-$\beta$ plasma, the dispersion relation
is $\omega=k V_{\A}$ (see Cho et al. 2002), and the required velocity for
TTD corresponds to $v_{\|}=v_{\A}/\cos\theta$. Thus, if $v_{\|}\ge V_{\A}$,
TTD can be efficient to accelerate grains to large velocities.
In the NLT limit, the diffusion coefficient for TTD is given by (see Appendix C)
\begin{eqnarray}
D_{pp}(\mu,p)^{\rm TTD}&=&\frac{v\sqrt{\pi}\Omega^{2}(1-\mu^{2})m^{2}V_{\A}^{2}
M_{\A}^{2}}{2LR^{2}}\int_{1}^{k_{c}L}x^{-5/2}dx\nonumber\\
&&\times\int_{0}^{1}\frac{d\eta}{\eta\Delta \mu}J_{1}^{2}(w){\rm exp}\left[-\frac{(\mu-\frac{V_{\A}}
{\eta v})^{2}}{(\Delta \mu)^{2}}\right].~~~~~~\label{eq:dpt}
\end{eqnarray}
\begin{figure}
\includegraphics[width=0.45\textwidth]{f3.ps}
\caption{Diffusion coefficient $D_{pp}^{{\rm TTD}}$ as a function of the cosine
of the grain pitch angle $\mu$ for the limit of QLT (green lines) and NLT (red lines).
Two velocity values of super-Alfv\'{e}nic graphite grains of size $a=10^{-5}$ cm
are considered. $D_{pp}^{{\rm TTD}}$ drops sharply for $\mu<V_{\A}/v$ in the QLT,
but $D_{pp}^{{\rm TTD}}$ is finite as $\mu \rightarrow 0$ in the NLT as a result of
broadening of resonance conditions.}
\label{Dpp_mu_TTD}
\end{figure}
In Figure \ref{Dpp_mu_TTD} we present $D_{pp}^{{\rm TTD}}$ as a function of the
cosine of the grain pitch angle $\mu$ obtained using the QLT and NLT.
Two values of the grain velocity $v=1.5~V_{\A}$ and $3.5~V_{\A}$ are considered.
It can be seen that in the former case, $D_{pp}^{{\rm TTD}}$ increases with decreasing
$\mu$ until $\mu=v/V_{\A}$, and drops sharply to zero for $\mu<v/V_{\A}$
because the resonance condition (\ref{eq:resTTD}) is not satisfied.
In contrast, the broadening of the resonance condition in the latter case allows grains
with $\mu<V_{\A}/v$ to have resonant interactions with waves, resulting in finite
$D_{pp}^{{\rm TTD}}$ even at $\mu=0$. Therefore, TTD is usually neglected in
the QLT because the gyroresonance tends to accelerate grains
in the perpendicular direction to $\Bv_{0}$, resulting in $\mu=0$, for which
$D_{pp}^{{\rm TTD}}\rightarrow 0$. However, TTD can play an important role
in driving grain motion when the fluctuations of guiding center are taken
into account.
\section{Grain Velocities due to Resonance Acceleration}
\subsection{Grain dynamics}
Consider an ensemble of grains with the same mass $m$, moving in the uniform
magnetic field with their different pitch angles $\mu$.
From the equation of motion, $mdv/dt=-mv/t_{{\rm drag}}+R$, where $R$ is the
random force, we can obtain
\begin{eqnarray}
m\frac{d\langle v^{2}\rangle}{dt}=-\frac{m\langle v^{2}\rangle}{t_{{\rm drag}}}+A(v),
\label{eq:dv-dt}
\end{eqnarray}
where $\langle v^{2}\rangle$ is the grain velocity dispersion averaged over the
ensemble of grains, $A(v)$ is the rate of energy gain (see YL03).
When the scattering is less efficient than the acceleration,\footnote{The efficiency of scattering relative to acceleration is described by the ratio $p^{2}D_{\mu \mu}/D_{pp}=\cos\theta^{-2}\left({v\cos\theta}/{V_{\A}}+\mu\right)^{2}$
where $\cos\theta=k_{\|}/k$ with $\theta$ is the angle between $\kv$ and $\Bv$ (see Appendix D).
The scattering is negligible for grains with $v< V_{\A}$, but it becomes important
for $v \ge V_{\A}$ or when the grain is moving along the magnetic field, i.e., $v_{\|}\gg v_{\perp}$. } the cosine of the grain pitch angle $\mu=\cos\beta$ changes
slowly during acceleration, and $A(v)$ is given by
\begin{eqnarray}
A(v)=\frac{1}{4p^{2}}\frac{\partial}{\partial p}\left(vp^{2}D_{pp}(p,\mu)\right),\label{eq:Av}
\end{eqnarray}
where $D_{pp}(p,\mu)$ is the diffusion coefficient.
When the scattering is more efficient than the acceleration, the pitch angle
can be rapidly redistributed through pitch angle diffusion during the acceleration
(i.e., diffusion approximation). The rate of energy gain (Eq. \ref{eq:Av}) is then
determined by the averaged value $D_{p}$ of $D_{pp}(p,\mu)$ over the
isotropic distribution of $\mu$ (see Dung \& Schlickeiser 1990ab):
\begin{eqnarray}
D_{p}(p)=\frac{1}{2}\int_{-1}^{1}\left(D_{pp}(p,\mu)-\frac{D_{\mu p}^{2}(p,\mu)}{D_{\mu\mu}(p,\mu)}\right)d\mu.\label{eq:dp}
\end{eqnarray}
We consider in the present paper zero helicity turbulence with $D_{\mu p}=0$.
When $D_{pp}$ is known, we solve Equation (\ref{eq:dv-dt}) iteratively to get
convergent velocities.
Physical parameters for ISM conditions are shown in Table 1. MHD turbulence
is injected at a large outer scale $L$ with velocity dispersion $\delta V$.
The value $\delta V$ is chosen such that the turbulence is weak and sub-Alfv\'{e}nic.
Major damping processes are also listed in Table 1. For the CNM and WNM,
the dominant damping arises from the neutral-ion viscosity. The value of $k_{c}$
for these phases is adopted from YLD04. For the WIM, the turbulent damping
arises mainly from ion viscosity and collisionless damping. We calculate $k_{c}$
by equating the damping rate to the rate of turbulence cascade
(see Appendix C, also YL03; YLD04).
Figure \ref{spectrum} sketches possible acceleration mechanisms present
in MHD turbulence. Working scales in the inertial range, spanning from
the injection scale $k_{\min}$ to the cutoff scale $k_{c}$ of turbulence, are indicated.
Gyroresonance works only if $k_{\res}\le k_{c}$, and is dominant at small scales
$k\ge k_{\res}\sim r_{g}^{-1}$. The critical size $a_{\cri}$ for gyroresonance is
then obtained by solving the equation $k_{\res}=k_{c}$ for grain size.
\begin{figure}
\includegraphics[width=0.45\textwidth]{f4.ps}
\caption{A sketch of spectrum of MHD turbulence and acceleration mechanisms
for charged grains are shown with its corresponding scale. $k_{\min}\sim L^{-1}$ is
the injection scale, $l_{\mfp}$ is the grain mean free path, $k_{\res}\sim r_{g}^{-1}$ is the
gyroresonance scale, and $k_{c}$ is the damping cut-off of turbulence.
We assume that grains of interest are large enough so that $k_{\res}<k_{c}$
or Larmor period larger than eddy turnover time.}
\label{spectrum}
\end{figure}
Figure \ref{Av} shows the rate of energy gain (Eq. \ref{eq:Av}) as a function
of grain velocity $v $ for gyroresonance acceleration ($n=\pm 1$) and TTD
acceleration ($n=0$) in the CNM ({\it upper panel}) and WIM ({\it lower panel})
arising from fast modes in MHD turbulence. Here we assumed that the pitch
angle scattering is efficient. Due to the broadening of resonance condition,
gyroresonant interactions occur at lower velocities in the NLT than QLT
({\it solid lines}). It can also be seen that the gyroresonance acceleration
($n=1$) is dominant for $v<V_{\A}$, while TTD acceleration
($n=0$) becomes dominant for $v\ge V_{\A}$.
In both the CNM and WIM, as $v\rightarrow V_{\A}$, $A_{v}(n=1)$ becomes
smaller in the NLT than the QLT. As a result, we expect that gyroresonance
acceleration is less efficient in the former case. We are going to quantify such
a difference in the following.
\begin{figure}
\includegraphics[width=0.45\textwidth]{f5a.ps}
\includegraphics[width=0.45\textwidth]{f5b.ps}
\caption{Rate of energy gain as a function of the grain velocity for gyroresonance
acceleration ($n=\pm 1$) and transit time damping acceleration ($n=0$, TTD)
for a graphite grain of size $a=10^{-5}$ cm in the CNM ({\it upper}) and WIM
({\it lower}). Solid and dot lines denote results from NLT and QLT, respectively.
The Alfv\'{e}nic speed $V_{\A}$ is indicated. The gyroresonance acceleration ($n=1$)
is dominant for $v< V_{\A}$, and the TTD acceleration ($n=0$) takes over
when $v\ge V_{\A}$. The case of efficient pitch angle scattering is considered.}
\label{Av}
\end{figure}
\subsection{Gyroresonance in quasi-linear theory and nonlinear theory}
Grain velocities due to gyroresonance acceleration are obtained by solving
Equation (\ref{eq:dv-dt}) using the diffusion coefficients from Equations (\ref{eq:dpg})
and (\ref{eq:dp}). We assume that at the beginning the grain has low velocity,
so that the pitch angle scattering by TTD is negligible. The gyroresonance
increases rapidly $v_{\perp}$, and $\mu$ decreases to $\mu=0$.
So, we can assume $\mu=0$ for the gyroresonance acceleration.
Figure \ref{Vcomp} shows grain velocities obtained using the NLT and QLT
for the CNM, WNM and WIM. Both silicate and graphite grains are considered.
Grain velocities obtained using the NLT are generically smaller than those from
the QLT. But the difference is within $15 \%$. The smaller results in the NLT
arise from the fact that, when grain velocities approach $V_{\A}$, a fraction of
turbulence energy is spent to induce transit time acceleration (see also Fig. 5).
The sudden cutoffs (dotted lines) present in the CNM and WNM
correspond to the cutoff scales of turbulence due to collisional and collisionless
damping occurring at the critical size $a_{\cri}$ (see Sec. 4.1).
\begin{figure}
\includegraphics[width=0.45\textwidth]{f6a.ps}
\includegraphics[width=0.45\textwidth]{f6b.ps}
\caption{Grain velocities relative to gas as a function of grain size for graphite
and silicate grains in various phases of the ISM. Acceleration arising from
gyroresonant interactions of fast modes with grains are obtained using the
QLT ({\it dotted line}) and NLT ({\it solid line}). The difference in velocities
from the NLT and QLT is within 15$\%$. The shaded area indicates the range
of size in which the assumption of constant charge is invalid due to strong
charge fluctuations. Dotted vertical lines present in the WNM and CNM denote
the critical size $a_{\cri}$ corresponding to the cutoff scale $k_{c}$ of turbulence.}
\label{Vcomp}
\end{figure}
Figures \ref{VNQcnm} and \ref{VNQwim} compare grain velocities arising from
fast MHD modes (similar data as Fig. \ref{Vcomp}) with those induced by
Alfv\'{e}n hydro-drag modes (LY02) and fast hydro-drag modes (YLD04)
in the CNM and WIM, respectively.
\footnote{Here only the velocity component perpendicular to magnetic field
is shown.}As expected from earlier studies, gyroresonance acceleration is
dominant for the entire range of grain size in the WIM (Fig. VNQwim).
The rapid variation of grain mean charge \Zmeant present in the WIM
(see Fig. \ref{sigmaZ}) results in the non monotonic
increase of grain velocities from gyroresonance (solid and dotted lines).
Two local maxima in the hydro drag cases correspond to the change in
sign of \Zmeant (see Fig. \ref{sigmaZ}). In the CNM, gyroresonance is
dominant for grain size from $a_{\rm cri}=5\times10^{-6}~\cm$ to
$a\sim 6\times 10^{-5}~\cm$, while acceleration by hydro-drag takes over
for grains smaller than $a_{\cri}$ and larger than $\sim6\times 10^{-5}~\cm$
(see Fig. \ref{VNQcnm}).
\begin{figure}
\includegraphics[width=0.45\textwidth]{f7a.ps}
\includegraphics[width=0.45\textwidth]{f7b.ps}
\caption{Comparison of grain velocities arising from gyroresonant interactions
of fast MHD modes ({\it solid and dotted lines}) with the results arising from
hydrodrag by fast and Alfv\'{e}n modes ({\it dot-dashed and dashed lines}).}
\label{VNQcnm}
\end{figure}
\begin{figure}
\includegraphics[width=0.45\textwidth]{f8a.ps}
\includegraphics[width=0.45\textwidth]{f8b.ps}
\caption{Similar to Fig. \ref{VNQcnm} but for the WIM. The shaded area indicates
the range of grain size in which the assumption of constant charge is invalid. Local
maximum velocity for hydro acceleration present in the range $a<10^{-6}$ and
$a>10^{-5}$ cm arise from the change in sign of $\langle Z\rangle$. }
\label{VNQwim}
\end{figure}
\subsection{Acceleration by TTD}
Gyroresonant acceleration tends to drive grain motion in perpendicular direction
to the mean magnetic field (i.e., $\mu=0$). As discussed in Section 3.2, the
broadening of resonance condition in the NLT allows TTD
to operate even at $\mu=0$. In addition to acceleration, the scattering by TTD
can be important, which results in the deviation of $\mu$ from $\mu=0$
(see Yan \& Lazarian 2008). However, it is still uncertain how fast the
pitch angle scattering (both gyroresonance and TTD) by fast MHD modes
is compared to the acceleration. For simplicity, we consider the TTD acceleration
for two limiting cases of {\it efficient} scattering and {\it inefficient} scattering
in which the scattering is more and less efficient than the acceleration, respectively.
In the latter case, the scattering is assumed to be sufficient to alter the adiabatic
invariant of gyromotion.
In the presence of TTD, we take into account the diffusion coefficients
$D_{pp}^{\rm TTD}$ from Equation (\ref{eq:dpt}) for Equation (\ref{eq:dv-dt})
in addition to $D_{pp}^{\rm G}$.
When the pitch angle scattering is efficient, $\mu$ is described by an isotropic
distribution $f(\mu)d\mu=1/2d\mu$. The diffusion coefficient $D_{p}(p,\mu)$
is replaced by its average value over the isotropic distribution. $f(\mu)$.
Figure \ref{VGT} compares grain velocities relative to gas obtained
using the NLT from gyroresonance ({\it dot line}), and
gyroresonance plus TTD by fast MHD modes for silicate grains ({\it upper})
and graphite grains ({\it lower}) assuming the efficient pitch angle scattering.
In the WIM conditions, the TTD acceleration is negligible for grains smaller than
$\sim 4\times 10^{-6}\cm$ for which $v<V_{\A}$. The efficiency of TTD begins
to increase rapidly with $a$ when $v\sim V_{\A}$. TTD can increase grain
velocities to an order of magnitude higher than gyroresonance.
The effect of TTD is less important in the WNM than the WIM, but still
considerable. For the CNM, TTD acceleration is rather marginal
because grains are moving with sub-Alfv\'{e}nic velocities, less than the
threshold for TTD.
\begin{figure}
\includegraphics[width=0.45\textwidth]{f9a.ps}
\includegraphics[width=0.45\textwidth]{f9b.ps}
\caption{ Grain velocities relative to gas arising from gyroresonant acceleration
({\it dotted line}) and gyroresonant acceleration plus transit time acceleration
({\it solid line}), as a function of grain size for different environments.
The case of efficient scattering is considered.}
\label{VGT}
\end{figure}
In the case of inefficient pitch angle scattering, the cosine of the grain
pitch angle $\mu=0$ is assumed as a result of gyroresonance acceleration.
We found that grain velocities are $\sim 10\%$ larger than the results for
the case of efficient scattering. This seems counterintuitive because
the scattering is required to alter the adiabatic invariant in gyromotion.
However, here we neglected that effect of pitch angle scattering, and
the situation is merely related to the fact that the diffusion coefficient
$D_{pp}$ averaged over $\mu$ is slightly lower than $D_{pp}$ at
$\mu=0$ (see Fig. 3).
\section{Stochastic acceleration by low frequency Alfv\'{e}n waves}
\subsection{General consideration}
The stochastic acceleration by low frequency Alfv\'{e}n waves with
$\omega < \Omega$ at the gyro-scale $k_{\perp}\sim \rho^{-1}$ was studied
in Chandran et al. (2010) for ion heating in the solar wind. Here we consider
the effect of low frequency Alfv\'{e}n waves on dust acceleration in the ISM.
For the low frequency Alfv\'{e}n waves, resonance acceleration is inefficient
because the resonance condition $\omega-k_{\|}v\mu=n\Omega$
is not satisfied. Indeed, in low-$\beta$ plasma, the ion thermal velocity
$v_{T}\ll V_{\A}$, and $\omega=k_{\|} V_{\A}$ for Alfv\'{e}n waves,
we have $\omega-k_{\|}v\mu\ll 0$.
Let $\delta v$ and $\delta B$ be the rms amplitudes of velocity and magnetic
field at the gyro-scale $k_{\perp}\rho\sim 1$. The electric field induced by
plasma perturbations with velocity $\delta v$ in the direction perpendicular
to the mean magnetic field has the rms amplitude
\begin{eqnarray}
\delta E\simeq\frac{\delta v B_{0}}{c},\label{eq:dE}
\end{eqnarray}
and the potential is written as
\begin{eqnarray}
\delta \Phi\simeq \rho \delta E.\label{eq:dphi}
\end{eqnarray}
The electric field in the plane perpendicular to $\Bv_{0}$ results in the
acceleration of grains in the perpendicular direction.
Combining these above equations, we obtain
\begin{eqnarray}
q\delta\Phi\simeq q\frac{\rho \delta vB_{0}}{c}=mv_{\perp}\delta v,\label{eq:dPhi}
\end{eqnarray}
where $m$ is mass of the charged particle and $\rho=mcv_{\perp}/qB_{0}$
is the gyro radius.
The fractional increase of energy after a single gyro-period is calculated by
\begin{eqnarray}
\frac{q\delta \Phi}{mv_{\perp}^{2}/2}\simeq 2\epsilon,
\end{eqnarray}
where
\begin{eqnarray}
\epsilon=\frac{\delta v}{v_{\perp}},\label{eq:epsilon}
\end{eqnarray}
When $\epsilon \ll 1$, also corresponding to $\delta B\ll B_{0}$ due to
the assumption of low $\beta$ plasma, the increase of energy per
gyroperiod is negligible because of adiabatic invariant for the magnetic
moment $mv_{\perp}^{2}/2B_{0}$. As $\epsilon$ increases to unity, the
particle energy changes substantially after one gyroperiod. The guiding
center becomes chaotic when $\epsilon$ exceeds some threshold value,
so that the perpendicular acceleration becomes important.
The increase of energy depends only on the amplitude of perturbation at
the scale of gyroradius. The increase of energy per unit of time per unit
of mass is defined as
\begin{eqnarray}
Q_{\perp}=\frac{v_{\perp}^{2}}{\tau_{\rm acc}},\label{Qperp1}
\end{eqnarray}
where $\tau_{\rm acc}$ is the time for the particle kinetic energy $K_{\perp}$
to increases by a factor of $2$.
We can estimate $\tau_{\rm acc}$ using the diffusion coefficient $D_{K}$
as follows:
\begin{eqnarray}
\tau_{\rm acc}=\frac{4K_{\perp}^{2}}{D_{K}}=\frac{m^{2}v_{\perp}^{4}}{D_{K}},\label{taui}
\end{eqnarray}
where
\begin{eqnarray}
D_{K}=\frac{\left(\Delta K_{\perp}\right)^{2}}{\Delta t}=m^{2}v_{\perp}^{2}\omega_{\eff}^{2}\delta v\rho,\label{DK}
\end{eqnarray}
and $\omega_{\eff}=\delta v/\rho$ is the effective frequency of gyroscale
fluctuations. Here we have taken $\Delta K_{\perp}\simeq mv_{\perp}\delta
v=mv_{\perp}\omega_{\eff}\rho$ from Equation (\ref{eq:dPhi}),
and $\Delta t=\rho/\delta v$ is the time required for the guiding center to
move by a distance equal to the gyro radius $\rho$.
Taking use of Equations (\ref{taui}) and (\ref{DK}) for (\ref{Qperp1}), we obtain
\begin{eqnarray}
Q_{\perp}\simeq \omega_{\eff}^{2}\delta v\rho\simeq\frac{\delta v^{3}}{\rho}.\label{Qperp2}
\end{eqnarray}
When $\epsilon$ is sufficiently small, the variation in the particle energy
is correlated over long time, so the assumption for diffusive approximation
may not be adequate. Thus, the actual energy gain is substantially smaller
than the value obtained by Equation (\ref{Qperp2}). To account for this
reduction, a damping function ${\rm exp}\left(-c_{2}/\epsilon\right)$ is
introduced, and Equation (\ref{Qperp2}) can be rewritten as
\begin{eqnarray}
Q_{\perp}=\frac{c_{1}(\delta v)^{3}}{\rho}{\rm exp}\left(-\frac{c_{2}}
{\epsilon}\right),\label{eq:qper}
\end{eqnarray}
where $c_{1}$ and $c_{2}$ are dimensionless constants that depend on the
nature of fluctuations. Below we assume $c_{1}=0.75$ and $c_{2}=0.34$ for
the fluctuations by low frequency Alfv\'{e}n waves as in Chandran et al. (2010).
\subsection{Grain velocities for the ISM}
Let assume that the Alfv\'{e}nic turbulence in the ISM follows the scaling
\begin{eqnarray}
\delta v_{\perp}=\alpha V_{\A}\left(\frac{\l_{\perp}}{L}\right)^{a}\label{eq:deltav}
\end{eqnarray}
where $l_{\perp}\sim \rho_{{\rm d}}$ and $L$ are gyro scale and the injection scale.
$\alpha$ and $a=(c_{3}-1)/2$ with $c_{3}$ being the slope of turbulence
power spectrum are dimensionless, derived from the properties of turbulence.
For sub-Alfv\'{e}nic turbulence, $\alpha<1$. Using Equation (\ref{eq:epsilon})
for dust grains, i.e., $\epsilon_{{\rm d}}=\delta v_{\perp}/v_{\perp,d}$, we obtain
\begin{eqnarray}
\epsilon_{d}=\alpha \left(\frac{B^{2}}{8\pi n_{\rm H}k_{\rm B}T_{\perp}} \right)^{(1-a)/2}\frac{A^{(1+a)/2}}{Z^{a}}\left(\frac{d_{p}}{L}\right)^{a},\label{eq:epschao}
\end{eqnarray}
where $p$ denotes proton, and $d$ denotes dust, $A=m/m_{p}$,
$d_{p}=V_{\A}/\Omega_{p}$, and the perpendicular temperature is defined as
$kT_{\perp}=m v_{\perp,{\rm d}}^{2}/2$.
Using Equations (\ref{eq:deltav}) and (\ref{eq:epschao}) for Equation (\ref{eq:qper}),
we obtain the rate of energy gain per a grain of mass $m$
\begin{eqnarray}
A(v_{\perp})=m\frac{c_{1}V_{\A}^{3}}{L}{\rm exp}\left(-\frac{c_{2}}
{\epsilon_{d}}\right).\label{eq:qper1}
\end{eqnarray}
For calculations, we assume $c_{3}=5/3$ and $a=1/3$ for Alfv\'{e}nic
turbulence above the gyro scale.
Using the parameters for the ISM in Table 1, we calculate the grain velocity
arising from the chaotic acceleration by low frequency Alfv\'{e}n waves in
Figure \ref{f4} for the CNM and WIM. We show that the chaotic acceleration
by low frequency Alfv\'{e}n waves is subdominant to the fast and Alfv\'{e}nic
hydrodynamic drag. Obviously, it is much less important than gyroresonance
and TTD by fast modes. The possible reason is that the low frequency
Alfv\'{e}n waves cascade faster to small scale than the fast modes.
\begin{figure}
\includegraphics[width=0.45\textwidth]{f10a.ps}
\includegraphics[width=0.45\textwidth]{f10b.ps}
\caption{Grain velocity due to stochastic acceleration by low frequency
Alfv\'{e}n waves (solid lines) compared to the acceleration by hydrodynamic
drag from Alfv\'{e}n modes (dotted lines) and fast modes (dashed lines),
for the CNM ({\it upper}) and WIM ({\it lower}). The stochastic acceleration
is much less efficient than the latter.}
\label{f4}
\end{figure}
\section{Discussion}
\subsection{Related works on dust grain acceleration}
The acceleration of dust grains by incompressible MHD turbulence was first studied
by Lazarian \& Yan (2002). Yan \& Lazarian (2003) studied grain acceleration
in compressible MHD turbulence, and discovered a new acceleration mechanism
based on gyroresonant interactions of grains with waves. This acceleration
mechanism increases grain velocities in perpendicular direction to the mean
magnetic field. YLD04 computed grain velocities arising from gyroresonance
by fast MHD modes using quasi-linear theory (QLT), and compared the obtained
results with different mechanisms, for various ISM phases. They found that the
gyroresonance is the most efficient mechanism for grain acceleration in the ISM.
The effect of large scale compression on grain acceleration is shown
by Yan (2009) to be less important than the gyroresonance in the ISM
conditions, unless the grains move with super-Alfv\'{e}nic velocities.
For very small grains (e.g., polycyclic aromatic hydrocarbons and nanoparticles),
Ivlev et al. (2010) sketched a new mechanism of grain acceleration due to
electrostatic interactions of grains with fluctuating charge and provided
rough estimates of grain velocities in the ISM. Hoang \& Lazarian (2011)
quantified this mechanism using Monte Carlo simulations of charge
fluctuations. They found that charge fluctuations can accelerate grains to
several times their thermal velocities.
\subsection{NLT for gyroresonance acceleration}
We have revisited the treatment of gyroresonance acceleration for charged
grains due to MHD turbulence by accounting for the fluctuations of grain guiding
center from a regular trajectory along the mean magnetic field (i.e.
NLT limit). The fluctuations of the guiding center result in the broadening of
resonance conditions-- a Delta function is replaced by a Gaussian function.
Such broadening of resonance condition allows some fraction of wave energy
spent through the TTD acceleration. As a result, grain velocities due to
gyroresonance acceleration are in general decreased by $\sim 15\%$
in the NLT limit.
\subsection{Transit time damping acceleration}
TTD acceleration is believed to be important when the parallel component of
grain velocity along the magnetic field exceeds the Alfv\'{e}n speed $V_{\A}$.
Although gyroresonance acceleration by fast modes can accelerate grains to
$v\ge V_{\A}$, their resulting velocity mostly perpendicular to the magnetic field,
i.e. $\mu=0$, makes TTD unfavored because the resonance condition
$\delta (\omega-k_{\|}v_{\|})$ is not satisfied. Indeed, we found that TTD is
efficient for $\mu>V_{\A}/v$ and negligible for $\mu<V_{\A}/v$ in the QLT limit.
This feature is consistent with the result for acceleration of cosmic rays in
Schlickeiser \& Miller (1998).
The situation changes when the fluctuations of the guiding center are taken
into account in the NLT. For this case, the resonance condition is broadened
beyond the $\delta$ function, and can be described by a Gaussian function.
As a result, TTD acceleration becomes important for $\mu <V_{\A}/v$,
including $90^{\circ}$ pitch angle.
In addition to acceleration, TTD also induces the grain pitch angle scattering,
which is dominant over the scattering by gyroresonance. Since the efficiency
of the TTD scattering is uncertain, we considered in the paper two limiting
cases of inefficient and efficient scattering in which the scattering is less and
more efficient than the acceleration. The pitch angle is equal to $90^{\circ}$
in the former, and isotropic in the latter.
When the scattering is more efficient than the acceleration, we showed that
for the WNM and WIM, the TTD acceleration can increase substantially the
grain velocity compared to results arising from gyroresonance.
Particularly, for grains larger than $5\times 10^{-6}$ cm in the WIM, TTD
acceleration is an order of magnitude greater than the gyroresonance
acceleration. TTD is clearly more efficient than the betatron acceleration studied
in Yan (2009). In the CNM, TTD acceleration is limited because the gyroresonance
acceleration is not able to speed up grains to super Alfv\'{e}nic stage.
When the scattering is less efficient than the acceleration, grain velocities are
within $10\%$ lower than the results for the efficient scattering case.
\subsection{Stochastic acceleration}
We study also the effect of low frequency ($\omega< \Omega$) Alfv\'{e}n
waves on dust grains in the ISM. We show that the stochastic acceleration
by low frequency Alfv\'{e}n waves is subdominant to the gyro-resonance
acceleration and TTD acceleration by fast modes. This may arise from the
fact that low frequency Alfv\'{e}n waves cascade faster to small scale than
the fast modes in MHD turbulence.
\subsection{Implication to dust coagulation and shattering and alignment}
Hirashita \& Yan (2009) adopted grain velocity due to gyroresonance
acceleration from YLD04 to model grain size distribution in the different ISM
conditions. Hirashita et al. (2010) studied grain coagulation and shattering in
the WIM for large dust grains ejected from Type II supernova. They showed
that the shattering of large dust grains due to turbulence plays an important
role in producing small size population that modifies extinction curves in
starburst galaxies.
The threshold velocity for the grain shattering is a function of the grain size:
\begin{eqnarray}
v_{\rm shat}=2.7\left(\frac{a}{10^{-7}~\cm}\right)^{-5/6} {\rm km}~\s^{-1},\label{v_cri}
\end{eqnarray}
where $v_{\rm dd}$ is the relative velocity of dust grains (Chokshi et al. 1993).
If $v_{\rm dd}< v_{\rm shat}$, the grains collide and stick together. When
$v_{\rm dd} > v_{\rm shat}$, the collisions with high velocity produce shock
waves inside the grains, and shatter them in smaller fragments. For
$v_{\rm dd}\geq 20$ km/s, the evaporation of the dust grain occurs and the
grains are destroyed.
When TTD is accounted for, grains larger than $5\times 10^{-6}$ in the WIM
and WNM may undergo efficient shattering because $v_{\rm dd}>V_{\A}=20$
km $\s^{-1}$.
The effects of high velocities we obtain on dust grain alignment require further
studies. Recent research has shown that the classical grain alignment theory
of Davis \& Greenstein (1951) (see also Lazarian 1995; Roberge \& Lazarian
1999 for more recent quantitative studies of the process) is subdominant to the
radiative torque (RAT) model (see Dolginov \& Mitrofanov 1976; Draine \&
Weingartner 1996; Lazarian \& Hoang 2007a). This model, however, was
criticized in Jordan \& Weingartner (2009) who appealed to the results of
gyroresonance acceleration of grains in Yan \& Lazarian (2003) and claimed
this means that fast moving grains will be randomized as their charge fluctuates.
Our results indicate that grains can be accelerated to even faster velocities,
which could make the problem for the alignment more severe. However, we
believe that the claim about suppression of the RAT alignment for fast moving
grains is a result of the confusion on the nature of the RAT alignment.
We plan to address this issue elsewhere. At the same time
high velocities of grains may induce another mechanical alignment of irregular
grains as it described in Lazarian \& Hoang (2007b). This alignment does
require further studies.
\section{Summary}
In the present paper, we study the resonance acceleration of charged grains
by fast modes and stochastic acceleration by Alfv\'{e}n waves in MHD turbulence.
Our main results are summarized as follows.
1. We revisit the treatment of gyroresonance acceleration of charged grains
in compressible MHD turbulence by taking into account the fluctuations of
grain guiding center from the regular trajectory along the mean field.
Gyroresonance interactions by fast modes can accelerate large grains to
super-Alfv\'{e}nic speed. We found that grain velocities are lower
by $15\%$ in the NLT than the QLT.
2. We investigate the effect of transit time damping (TTD) by fast modes
for super-Alfv\'{e}nic grains. We found that the fluctuations of grain guiding
center allow TTD to occur not only within the range of the cosine of the grain
pitch angle $\mu>V_{\A}/v$ as expected by the QLT, but also for $\mu<V_{\A}/v$.
We show that the TTD acceleration can increase grain velocities by an order of
magnitude compared to the results arising from gyroresonance mechanism
Thus, TTD is the most efficient acceleration mechanism for super-Alfv\'{e}nic grains.
3. The stochastic acceleration due to low frequency Alfv\'{e}n waves is
inefficient for dust grains in the ISM conditions.
\acknowledgements
AL thanks Alexander von Humboldt Foundation. R.S. acknowledges the
support from the Deutsche Forschungsgemeinschaft through grants Schl
201/19-1 and Schl 201/23-1. TH and AL acknowledge the support of the Center for
Magnetic Self-Organization. We thank Bruce Draine for providing us the data
of grain charge distribution and Huirong Yan for valuable comments.
We thank the anonymous referee for her/his useful comments that improve the paper.
|
\section{Introduction}
The cosmological dark matter can
consist of dark atoms, in which new stable charged particles are bound by ordinary Coulomb interaction (See \cite{Levels,Levels1,mpla} for review and references).
In order to avoid anomalous
isotopes overproduction, stable particles with charge -1 (and
corresponding antiparticles), as tera-particles \cite{Glashow}, should be absent \cite{Fargion:2005xz}, so that stable
negatively charged particles should have charge -2 only.
Such stable double charged particles can hardly find place in SUSY models, but there
exist several alternative elementary particle frames, in which heavy
stable -2 charged species, $O^{--}$, are predicted:
\begin{itemize}
\item[(a)] AC-leptons, predicted
in the extension of standard model, based on the approach
of almost-commutative geometry \cite{Khlopov:2006dk,5,FKS,bookAC}.
\item[(b)] Technileptons and
anti-technibaryons in the framework of walking technicolor
models (WTC) \cite{KK,Sannino:2004qp}.
\item[(c)] and, finally, stable "heavy quark clusters" $\bar U \bar U \bar U$ formed by anti-$U$ quark of 4th
\cite{Khlopov:2006dk,Q,I,lom} or 5th \cite{Norma} generation.
\end{itemize}
All these models also
predict corresponding +2 charge particles. If these positively charged particles remain free in the early Universe,
they can recombine with ordinary electrons in anomalous helium, which is strongly constrained in the
terrestrial matter. Therefore cosmological scenario should provide a mechanism, which suppresses anomalous helium.
There are two possibilities, requiring two different mechanisms of such suppression:
\begin{itemize}
\item[(i)] The abundance of anomalous helium in the Galaxy may be significant, but in the terrestrial matter
there exists a recombination mechanism suppressing this abundance below experimental upper limits \cite{Khlopov:2006dk,FKS}.
\item[(ii)] Free positively charged particles are already suppressed in the early Universe and the abundance
of anomalous helium in the Galaxy is negligible \cite{mpla,I}.
\end{itemize}
These two possibilities correspond to two different cosmological scenarios of dark atoms. The first one is
realized in the scenario with AC leptons, forming neutral AC atoms \cite{FKS}.
The second assumes charge asymmetric case with the excess of $O^{--}$, which form atom-like states with
primordial helium \cite{mpla,I}.
If new stable species belong to non-trivial representations of
electroweak SU(2) group, sphaleron transitions at high temperatures
can provide the relationship between baryon asymmetry and excess of
-2 charge stable species, as it was demonstrated in the case of WTC
\cite{KK,KK2,unesco,iwara}.
After it is formed
in the Standard Big Bang Nucleosynthesis (SBBN), $^4He$ screens the
$O^{--}$ charged particles in composite $(^4He^{++}O^{--})$ {\it
O-helium} ``atoms'' \cite{I}.
In all the proposed forms of O-helium, $O^{--}$ behaves either as lepton or
as specific "heavy quark cluster" with strongly suppressed hadronic
interaction. Therefore O-helium interaction with matter is
determined by nuclear interaction of $He$. These neutral primordial
nuclear interacting objects contribute to the modern dark matter
density and play the role of a nontrivial form of strongly
interacting dark matter \cite{McGuire:2001qj,Starkman}.
The cosmological scenario of O-helium Universe allows to explain many results of experimental searches for dark matter \cite{mpla}. Such scenario is insensitive to the properties of $O^{--}$, since the main features of OHe dark atoms are determined by their nuclear interacting helium shell. It challenges direct experimental search for the stable charged particles at accelerators and such search strongly depends on the nature of $O^{--}$.
Stable $-2$ charge states ($O^{--}$) can be elementary like AC-leptons or technileptons,
or look like elementary as technibaryons. The latter, composed of techniquarks, reveal their structure at much higher energy scale and should be produced at LHC as
elementary species. They can also be composite like "heavy quark
clusters" $\bar U \bar U \bar U$ formed by anti-$U$ quark in one of the models of fourth
generation \cite{Q,I} or $\bar u_5 \bar u_5 \bar
u_5$ of (anti)quarks $\bar u_5$ of stable 5th family in the approach
\cite{Norma}.
In the context of composite dark matter scenario accelerator search for stable particles
acquires the meaning of critical test for existence of
charged constituents of cosmological dark matter.
The signature for AC leptons and techniparticles is unique and distinctive what allows
to separate them from other hypothetical exotic particles.
In particular, the ATLAS
detector has an unique potential to identify these particles and measure their masses.
Test for composite $O^{--}$ can be only indirect:
through the search for heavy hadrons, composed of single $U$ or
$\bar U$ and light quarks (similar to R-hadrons).
Here we
study a possibility for experimental probe of this hypothesis.
\section{\label{quarks} New stable generations}
Modern precision data
on the parameters of the Standard model do not exclude \cite{Maltoni:1999ta} the existence of
the 4th generation of quarks and leptons.
In one of the approaches the 4th generation follows from heterotic string phenomenology and
its difference from the three known light generations can be
explained by a new conserved charge, possessed only by
its quarks and leptons \cite{Q,I,lom,Belotsky:2000ra}. Strict conservation of this charge makes the
lightest particle of 4th family (neutrino) absolutely
stable, but it was shown in \cite{Belotsky:2000ra} that this neutrino cannot be the dominant form of the dark matter.
The same conservation law requires the lightest quark to be long living
\cite{Q,I}. In principle the lifetime of $U$ can exceed the age of the
Universe, if $m_U<m_D$ \cite{Q,I}.
In the current implementation of the "{\it spin-charge-family-theory}" \cite{Norma} there are predicted two sets with four generations each, so that the 4th generation is unstable, while the lightest (5th generation) of the heavy set has no mixing with light families and thus is stable. If $m_{u_5}<m_{d_5}$ and their mass difference is significant, OHe dark matter cosmological scenario can be realized in this theory. For the lower possible mass scale ($\sim 1 TeV$) for the 5th generation particles, their search at LHC is possible along the same line as for stable particles of 4th generation in the approach \cite{Q,I,lom,Belotsky:2000ra}. In the successive discussion we'll consider stable $u$-type quark without discrimination of the cases of 4th and 5th generation, denoting the stable quark by $U$.
Due to their Coulomb-like QCD attraction ($\propto \alpha_{c}^2 \cdot m_U$, where $\alpha_{c}$ is the QCD constant) stable double and triple $U$ bound states $(UUq)$, $(UUU)$ can exist
\cite{Q,Glashow,Fargion:2005xz,I,lom,Norma}. The corresponding antiparticles can be formed by heavy antiquark $\bar U$. Formation of these double and
triple states at accelerators and in
cosmic rays is strongly suppressed by phase space constraints, but they can be formed in early
Universe and strongly influence cosmological evolution of 4th
generation hadrons. As shown in \cite{I}, \underline{an}ti-
\underline{U}-\underline{t}riple state called \underline{anut}ium
or $\Delta^{--}_{3 \bar U}$ is of a special interest. This stable
anti-$\Delta$-isobar, composed of $\bar U$ antiquarks can be bound with $^4He$ in atom-like state
of O-helium \cite{Khlopov:2006dk}.
Since simultaneous production of three $U \bar U$ pairs and
their conversion in two doubly charged quark clusters $UUU$
is suppressed, the only possibility to test the
models of composite dark matter from 4th (or 5th) generation in the collider experiments is a search for production of stable hadrons containing single $U$ or $\bar U$.
$U$-quark can form lightest $(Uud)$ baryon and $(U \bar u)$ meson
with light
quarks and antiquarks. $\bar U$ can form the corresponding stable antiparticles, like $\bar U \bar u \bar d$ and $\bar U u$. Search for these stable hadrons is similar to the R-hadrons search. The main task will be to distinguish R-hadrons from hadrons, containing quarks of 4th or 5th generation. R-hadrons will be accompanied by supersymmetric particles, what is not the case for 4th or 5th generation hadrons.
\section{\label{accelerators} Signatures for $U$-hadrons in accelerator experiments}
In spite of that the mass of $U$-quarks can be quite close to that of $t$-quark, strategy of their search should be completely different. $U$-quark in framework of the considered models is stable and will form stable hadrons at accelerator contrary to $t$-quark.
Detailed analysis of possibility of $U$-quark search requires quite deep understanding
of physics of interaction between (meta-)stable U-hadrons and nucleons of matter.
However, methodic for $U$-quark search can be described in general, if we
know mass spectrum of $U$-hadrons and (differential) cross sections of their production.
Cross section of $U$-quark production in pp-collisions
is presented on the Fig. \ref{crossec}. For comparison, cross sections of 4th generation
leptons are shown too. Cross sections have been calculated with program CompHEP \cite{CompHEP}. Cross sections of $U$- and $D$- quarks virtually do not differ.
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{Xsection.eps}
\end{center}
\caption{Cross sections of production of 4th generation particles (N, E, U (D)) at LHC. Solid and dashed curves correspond to c.m. energies 7 and 14 TeV respectively.
Horizontal dashed line shows approximate level of sensitivity to be reached in 2012 (at the energy 7 TeV).}
\label{crossec}
\end{figure}
For quarks ($U$ and $D$) the obtained values were re-scaled in correspondence of estimations done with program Hathor \cite{Hathor}.
Heavy stable quarks will be produced with high transverse
momentum $p_T$ and velocity, which is less than speed of light. In general,
simultaneous measurement of velocity and momentum provides us
information about mass of particle. Information on ionization
losses are, as a rule, not so good. All these features are
typical for any heavy particle, while there can be subtle
differences in the shapes of their angle- and $p_T$-distribution,
defined by concrete model, which it predicts. It is the peculiarity of
long-lived hadronic nature what can be of special importance for
clean selection of events of $U$-quarks production.
$U$-quark can form a
whole class of $U$-hadron states which can be considered as stable
in the conditions of an accelerator experiment contrary to their relics in Universe.
But in any case, as we pointed out, double and triple $U$-hadronic states cannot
be virtually created at collider. Many other hadronic states, whose
lifetime exceeds $\sim 10^{-7} \s$, should also look as stable in accelerator experiment. In the
Table 1 expected mass spectrum of $U$-hadrons, obtained with the help of
code PYTHIA \cite{pythia}, is presented.
\begin{table}
\caption{Mass spectrum and relative yields in LHC for U-hadrons}
\begin{tabular}{|p{1.3in}|p{1.1in}|p{1.2in}|p{0.55in}|} \hline
& {\small Difference of masses of U-hadron and U-quark, GeV} & \multicolumn{2}{|p{1.9in}|}{\small Expected yields in \% (in the right columns the yields of long-lived stated are given)} \\ \hline
$\left\{U\tilde{u}\right\}^{0} ,\left\{U\tilde{d}\right\}^{+} $ & 0,330 & \multicolumn{2}{|p{1.8in}|}{39,5(3)\%, 39,7(3)\%} \\ \hline
$\left\{U\tilde{s}\right\}^{+} $ & 0,500 & \multicolumn{2}{|p{1.8in}|}{11,6(2)\%} \\ \hline
$\left\{Uud\right\}^{+} $ & 0,579 & 5,3(1)\% & 7,7(1)\% \\ \hline
$\left\{Uuu\right\}_{1}^{++} ,\left\{Uud\right\}_{1}^{+} ,$ $\left\{Udd\right\}_{1}^{0} $ & 0,771 & 0,76(4)\%, 0,86(5)\%, 0,79(4)\% & \\ \hline
$\left\{Usu\right\}^{+} ,\left\{Usd\right\}^{0} $ & 0,805 & 0,65(4)\%, 0,65(4)\% & 1,51(6)\% \\ \hline
$\left\{Usu\right\}_{1}^{+} ,\left\{Usd\right\}_{1}^{0} $ & 0,930 & 0,09(2)\%, 0,12(2)\% & \\ \hline
$\left\{Uss\right\}_{1}^{0} $ & 1,098 & \multicolumn{2}{|p{1.8in}|}{0,005(4)\%} \\ \hline
\end{tabular}
\end{table}
The lower indexes in notation of $U$-hadrons in the Table 1 denote the nonzero spin $s=1$ of the pair of light quarks.
From comparison of masses of different $U$-hadrons
it follows that all $s=1$ $U$-hadrons decay quickly emitting $\pi$-meson or $\gamma$-quantum,
except for $(Uss)$-state. In the right column the expected relative yields are presented.
Unstable $s=1$ $U$-hadrons decay onto
respective $s=0$ states, increasing their yields.
There are two
mesonic states being quasi-degenerated in mass: $U\bar u$ and $U\bar d$
(we skip here discussion of strange $U$-hadrons).
Interaction with the medium composed of $u$ and
$d$ quarks transforms $U$-hadrons into those ones containing
$u$ and $d$ (as it is the case in the early Universe \cite{Q,I,lom}).
The created pair of $U\,\bar U$ quarks will fly
out of the vertex of pp-collision as $U$-hadrons with
positive charge in 60\% of all $U$-quark events
and as neutrals in 40\% (correspondingly, 60\% with negative charge and 40\% neutrals
for $\bar U$ hadrons).
After traveling through the matter of detectors, at a distance of a few nuclear lengths from
vertex, $U$-hadrons will transform in (roughly) 100\% of positively
charged hadrons $(Uud)$, while $\bar U$ -hadrons will convert in 50\% into
negatively charged $\bar U$ -hadron $(\bar U d)$ and in 50\% to neutral
$\bar U$ -hadron $(\bar U u)$.
This feature will enable to discriminate the considered case of $U$-quarks from variety of alternative models,
predicting new heavy stable particles.
Note that if the mass
of Higgs boson exceeds $2m$, its decay channel into the pair of stable
$Q \bar Q$ will dominate over the $t \bar t$, $2W$, $2Z$ and
invisible channel to neutrino pair of 4th generation
\cite{nuHiggs}. It may be important for the strategy of heavy
Higgs searches.
\section{Conclusions}
The cosmological dark matter can be formed by
stable heavy double charged particles bound in neutral OHe dark atoms with primordial He nuclei by ordinary Coulomb interaction. This scenario sheds new light on the nature of dark matter and offers nontrivial solution for the puzzles of direct dark matter searches. It can be realized in the model of stable 4th generation or in the approach unifying spin and charges and challenges for experimental probe at accelerators.
In the context of this scenario search for new heavy stable quarks acquires the meaning of direct experimental probe for charged constituents of dark atoms of dark matter.
The $O^{--}$ constituents of OHe in the model of stable 4th generation and in the "{\it spin-charge-family-theory}" are "heavy quark clusters" $\bar U \bar U \bar U$. Production of such clusters (and their antiparticles) at accelerators is virtually impossible. Therefore experimental test of the hypothesis of stable $U$ quark is reduced to the search for stable or metastable $U$ hadrons, containing only single heavy quark or antiquark. The first year of operation at the future 14 TeV energy of the LHC has good discovery potential for $U$($D$)-quarks with mass up to 1.5 TeV, while the level of sensitivity expected in the 2012 at the LHC energy 7 TeV can approach to the mass of 1 TeV. $U$-hadrons born at accelerator will
distinguish oneself by high $p_T$, low velocity, by effect of a charge flipping
during their propagation through the detectors. All these features
enable to strongly increase the efficiency of event selection from not only
background but also from alternative hypothesis. In particular, we show that the detection of positively charged $U$-baryon in coincidence with $\bar U$-mesons (50\% neutrals and 50\% negatively charged) provides a distinct signature for the stable $U$ quark. Analysis of other channels of new particles production provides distinctions from the case of R-hadrons. In the latter case all the set of supersymmetric particles should be produced.
It should be noted that the "{\it spin-charge-family-theory}" predicts together with stable 5th generation also 4th generation of quarks and leptons, which are mixed with the three known families and thus unstable. Experimental probe for new unstable heavy particles implies definite prediction for their mass spectrum and branching ratios for their modes of decay.
\section {Acknowledgments}
We would like to thank Norma Mankoc-Borstnik, all the
participants of Bled Workshop and A.S.Romaniouk for stimulating discussions.
|
\section*{\label{Intro_1}Previous comments:
on the relevance of all this in view of past and recent tests}
\begin{center} \textit{(arXiv version)} \end{center}
Before anything else,
the tests by Brida \textit{et al} \cite{Brida_et_al02} do not disprove
the core of the Wigner-PDC approach as developed in
\cite{pdc1,pdc2,pdc3,pdc4,pdc5,pdc6,pdc7};
it only does so with some detection models proposed to be able to interpret the
expressions in consistency with local-realism (LR).
Precisely what we prove here is that such (local-realist) interpretation does
not need to depart from QM at any step (at least for the relevant subset of quantum
mechanical states),
a property that former proposals like \cite{W_LHV_02} did not satisfy (not to
mention other problems of their own).
This said, neither is ours here the ultimate proof that everything in regard
to the Wigner-PDC approach works perfectly fine: well, I claim it is so, but
so far only at the mathematical level.
The one and only value of this work is to show that, even still
in need to determine further choices both at the physical and purely mathematical
levels, there is room to accommodate all necessary restrictions.
Once more, \textit{without any need to depart from quantum mechanical predictions,
provided we stay inside the subset of QED-states compatible with LR}.
It should also be clear that further extensions of this formalism, giving
rise to new predictions such as the ``spontaneous Parametric Up Conversion''
proposed in \cite{pdc10} (and apparently also disproved by experimental work
in \cite{BG03}), are not addressed (and hence are irrelevant for our discussions)
here.
Now,
I am still convinced that no conclusive proof of the violation of local-realism
(LR) has ever been obtained:
we have the absence of proper space-like separation or ``locality loophole''
in experiments with massive particles, and the so-called ``detection loophole''
in experiments with photons, and something related to this last in the latest
important development (\cite{Kot_et_al12}, see Secs.\ref{Comp}).
Each of these ``loopholes'' places the corresponding experimental observations
within the frontiers of LR, and once there nothing is happening that cannot
(and should not try to) be understood from a purely classical framework, or
at least from an intermediate one such as the one we deal with here.
I use ``intermediate'' in the sense that what we have is a quantum formalism,
but we are also incorporating, in an implicit way by considering the vacuum state
which admits a well defined joint probability density as follows from the
positivity of the Wigner function, the limits imposed by LR.
Indeed, those latest developments in \cite{Kot_et_al12} are, in
a way, even more compelling from the point of view of my claims than the usual
Bell experiments are:
while these last are usually interpreted as proof of the ``non-local'' nature
of QM, which is a way out to respect pure realism (``something is there even
if we do not look at it''),
results in \cite{Kot_et_al12} would suggest the non-existence of a well defined
joint probability density for the results of a set of measurements performed
locally upon a system, which is also that (the non-existence) of realism itself.
After all, non-localities, such as action-at-a-distance, are also present in
other branches of physics like for instance electromagnetism,
but then always reconciled with LR once the full problem is considered (in regard
to e.m., the apparition of retarded potentials), and at least showing reasonable
properties such a decrease with growing distance between the parties.
In view of \cite{Kot_et_al12}, it is not any longer a matter of just some
instantaneous interaction that does not fade away with distance, but something
clearly more fundamental:
``things are not there until we look at them''.
The widespread willingness to accept such a view of physical reality before
exhausting other alternatives seems demeaning to the concept of science itself:
those alternatives have been proposed \cite{pdc1,pdc2,pdc3,pdc4,pdc5,pdc6,pdc7},
including basic elements (a random background of vacuum field fluctuations)
that do not look at all alien from the perspective of classical electromagnetism,
and that even some recent works within the orthodox approach to the
field now acknowledge \cite{Steering12}.
Besides, and to my knowledge, \cite{Kot_et_al12} is the only amongst other
recent related tests \cite{Peruzzo12, Kaiser12} which does not acknowledge,
explicitly, one or other loophole leading to compatibility with the
local-realistic interpretation.
In any case, in absence (I claim) of conclusive results, perhaps it is time
to consider the possibility that not all states allowed by QM may have a
physical counterpart, in particular those that would yield correlations
defying LR.
Perhaps is also time to start giving credit to models that, even though
being quantum-mechanical, may include that restriction built within their
formulation; here I propose a step in this direction.
This step is necessary, but of course not conclusive: what we treat here
is just that, a ``model''.
I hope it encourages further research, at least.
\section{\label{Intro}Introduction}
The Wigner picture of Quantum Optics of photon-entanglement generated by Parametric
Down Conversion \cite{PDC_explanation} was developed some years ago in a series of
papers \cite{pdc1,pdc2,pdc3,pdc4,pdc5,pdc6,pdc7};
recently, the approach has been revitalized producing another stream of very
interesting results \cite{pdc8,pdc9}.
It departs from an stochastic electrodynamical description (hence, based on
continuous variables: electromagnetic fields defined at each and every point of space),
and by use of the so-called Wigner transformation \cite{KimNoz} it
acquires a form where all expectation values depend on a certain probability
distribution for the value of those fields in the vacuum.
Such a picture clearly differs, on the other hand, from the usual one in Quantum
Information (QInf), based on a discrete description of the photon (a particle)
and not on a set of fields in a continuum space.
For instance, in a polarizing beam splitter (PBS), empty polarization channels
at either of the exits of the device are filled with the random components of
the vacuum Zero-Point field (ZPF); these random components can for instance
give rise to the enhancement of the detection probability for certain realizations
of the state of the fields (certain photons, in the QED language): see note
\cite{WignerPDC_enhancement}.
Phenomena such as this last come as a big difference with the customary model
of the set polarizer-detector is treated just as a ``black box'', able to extract
polarization information
in principle without the (explicit) intervention of any additional noise.
The Wigner-PDC framework provides an alternative, apparently local-realistic
explanation of the results of many typically quantum experiments,
the result of a measurement depending on (and only on) the set of hidden
variables (HV) inside its light cone:
in this case, both the signal propagating from the source and the additional noise
introduced by the ZPF at intermediate devices and detectors.
Within the first series of papers \cite{pdc1,pdc2,pdc3,pdc4,pdc5,pdc6,pdc7} those
experiments included:
frustrated two photon creation via interference \cite{pdc1};
induced coherence and indistinguishableness in two-photon interference
\cite{pdc1,pdc2};
Rarity and Tapster's 1990 experiment with phase-momentum entanglement
\cite{pdc2};
Franson's (original, 1989) experiment \cite{pdc2};
quantum \textit{dispersion cancellation} and Kwiat, Steinberg and Chiao's
\textit{quantum eraser} \cite{pdc3}.
From the most recent one \cite{pdc8,pdc9} we can also add,
on one side, amongst quantum cryptography experiments based on PDC:
two-qubit entanglement and cryptography \cite{pdc8} and quantum key distribution
and eavesdropping \cite{pdc8};
on the other, an interpretation for the experimental partial measurement of the
Bell states generated from a single degree of freedom (polarization) in \cite{pdc9}.
We have also had access to some promising (but still unpublished) work
on \textit{hyper-entanglement} \cite{Casado_priv_1}.
Perhaps needless to say, this "local realistic" picture of the quantum experiments
takes place within the limits (detection efficiencies) where it is already well
acknowledged that such an explanation may exist;
moreover, those limits arise as natural consequences of the theory: they are
simply limiting values of detection rates for states of light that are ultimately
generated from the vacuum.
This last feature is what makes the Wigner approach so interesting;
on the other hand, above we have used ``apparently'' because that local-realistic
interpretation, even within the corresponding limits of detection rates, was until
now not devoid of difficulties, in particular related to the detection model (so
far merely a one-to-one counterpart with Glauber's original expressions
\cite{detection}):
as a result of the normal order of operators, there average intensity due to vacuum
fluctuations is ``subtracted'', a subtraction that seems to introduce problems
related to the appearance of what could be interpreted as ``negative probabilities''.
That last is the issue that interests us here, one that, as to be expected,
did motivate the proposal of several modifications upon the expressions for
the detection probabilities:
for instance, as early as in \cite{pdc4}, ``our theory is also in almost perfect
one-to-one correspondence with the standard Hilbert-space theory, the only difference
being the modification in the detection probability that we proposed in relation...''.
Those modifications ranged from the mere inclusion of temporal and spatial
integration \cite{pdc4} to the proposal of much more complicated functional
dependencies \cite{pdc7,W_LHV_02,MS2008},
all of them seeming to pose their own problems; for instance in \cite{W_LHV_02},
a departure from the quantum predictions at low or high intensities,
experimentally disproved for instance in \cite{Brida_et_al02}.
Our route here is a different one however:
we do not propose any modification of the initial expressions for the detection
probabilities (in one-to-one correspondence with the initial quantum electrodynamical
model), but just explore the possibility of performing some convenient mathematical
manipulation that casts them in a form consistent with the axioms of probability,
hence one consistent with local-realism.
Following for instance \cite{pdc4} (see also \cite{detection}),
quantum mechanical detection probabilities, single and joint, can respectively
be expressed as
\begin{eqnarray}
P_{i}
&\propto& \langle I_i - I_{0,i} \rangle \nonumber\\
&=& \int_{\alpha,\alpha^{*}}
( I_{i}(\alpha,\alpha^{*}) - I_{0,i} )\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*},
\label{marginal}
\end{eqnarray}
\begin{eqnarray}
P_{i,j} &\propto&
\int_{\alpha,\alpha^{*}}
( I_{i}(\alpha,\alpha^{*}) - I_{0,i} ) \cdot ( I_{j}(\alpha,\alpha^{*}) - I_{0,j} )
\nonumber\\
&&\quad\quad\quad\quad\quad\quad\quad\quad\quad
\times\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*},
\label{joint}
\end{eqnarray}
where $\alpha,\alpha^{*}$ are vacuum amplitudes
of the (relevant set of) frequency modes \cite{note_modes} at the entrance
of the crystal, $W(\alpha,\alpha^{*})$ is obtained as the Wigner transform of the vacuum
state, $I_{i}(\alpha,\alpha^{*})$ is the field intensity (for that mode) and $I_{0,i}$ the
mean intensity due to the vacuum amplitudes, both at the entrance of the $i$-th detector
(see \cite{note_modes}):
\begin{eqnarray}
I_0 = \int_{\alpha,\alpha^{*}}
I_{0}(\alpha,\alpha^{*})\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*}.
\end{eqnarray}
The last three expressions would in principle allow us to identify the vacuum amplitudes
with a vector of hidden variables $\lambda \in \Lambda$ ($\Lambda$ is the space of events
or probabilistic space), with an associated density function $\rho(\lambda)$,
\begin{eqnarray}
\lambda &\equiv& \alpha,\alpha^{*}, \label{lambda}\\
\rho(\lambda) &\equiv& W(\alpha,\alpha^{*}). \label{rho_lambda}
\end{eqnarray}
Now, for instance in \cite{pdc4} it is already acknowledged that (\ref{marginal})--(\ref{joint})
cannot, because of the possible negativity of the difference $I_{i}(\alpha,\alpha^{*}) - I_{0,i}$,
be written as
\begin{eqnarray}
P_{i} & = & \int_{\Lambda}{ P_{i}(det|\lambda) \ \rho(\lambda)\ d\lambda }, \\
\label{lambda_single}
P_{i,j} & = &
\int_{\Lambda}{ P_{i,j}(det|\lambda)\ \rho(\lambda)\ d\lambda }, \nonumber\\
\label{lambda_joint}
\end{eqnarray}
where naturally $P_{i}(det|\lambda), P_{i,j}(det|\lambda)$ should stay positive (or
zero) always.
This last is our point of departure: in this paper we propose a reinterpretation of
the former marginal and joint detection probabilities based on a certain manipulation
of expressions (\ref{marginal})--(\ref{joint}).
The paper is organized as follows.
In Sec.\ref{No_pol} we will consider a setup with just a source and two detectors;
once that is understood, the interposition of other devices between the source and
the detectors poses no additional conceptual difficulty though it is nevertheless
convenient to address it in some detail: this will be done in Sec. \ref{Pol}.
The calculations in these two sections find support on the proofs provided in Appendix
\ref{Math}, and stand for our main result in this paper.
Sec. \ref{Fac} explores the question of ``$\alpha$-factorability'' in the model,
not only from the mathematical point of view but also providing some more physical
insights on its implications.
Up to that point the novel points of the paper are made and its results are
self-contained;
it is nevertheless natural to extend our analysis to some of their further
implications (amongst these, the consequences for Bell tests of local-realism,
and also a recent related proposal), in Sec. \ref{Comp},
where we also include a preliminary approach to questions regarding the physics of the
real detectors.
Finally, overall conclusions are presented in Sec. \ref{Conc}, and some
supplementary material is provided in Appendix \ref{Rev}, which may
not only help make the paper self-contained but perhaps also contribute to clarify
some of the questions addressed,
in particular the non-factorability issue.
\begin{figure}[ht!] \includegraphics[width=1.0 \columnwidth,clip]{WignerPDC_1.eps}
\caption{
Wigner-PDC scheme: photon pair generation, polarizing beam splitters (PBS) and detectors.
Only relevant inputs of Zero-Point vacuum field (ZPF) are represented in the picture:
``relevant'' can be understood, in a classical wavelike approach, as ``necessary
to satisfy energy-momentum conservation'' for the (set of) frequency modes of interest;
in a purely quantum electrodynamical one we would be talking about conservation of the
commutation relations at the empty exit channels of the devices.
Besides, those new ZPF components introduced at the empty exit channels of PBS's 1
and 2 can alter the detection probability for the signal arriving to the detectors
(for instance giving rise to its ``enhancement'');
that signal is on the other hand determined (amongst other hidden variables) by the
ZPF components entering the crystal.
Figure: courtesy of A. Casado.
} \label{Scheme} \end{figure}
\section{\label{No_pol}
Reinterpreting detection probabilities}
Our reinterpretation of expressions (\ref{marginal})--(\ref{joint}) involves two
steps:
\vspace{0.2cm}\noindent
(A)
Knowing that the following equality holds (from here on we drop detector indexes
when unnecessary), for some real constant $K_{(m)}$ (``marginal''),
\begin{eqnarray}
&&K_{(m)} \int_{\alpha,\alpha^{*}}
(I(\alpha,\alpha^{*}) - I_0)\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*}
\nonumber\\
&&\quad\quad\quad\quad\quad\quad\quad\quad
= \int_{\Lambda}{ P(det|\lambda) \ \rho(\lambda)\ d\lambda },
\label{eq_marginal}
\end{eqnarray}
we realize we do not need to assume
\begin{eqnarray}
P(det|\lambda) & \equiv & K_{(m)} \cdot (I(\alpha,\alpha^{*}) - I_0),
\end{eqnarray}
as a necessary, compulsory choice;
it would be enough to find some $f(x) \geq 0$, satisfying
\begin{eqnarray}
&& K_{(m)} \int_{\alpha,\alpha^{*}}
(I(\alpha,\alpha^{*}) - I_0)\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*}
\nonumber\\
&&\quad\quad\quad\quad\quad\quad
= \int_{\alpha,\alpha^{*}}
f(I(\alpha,\alpha^{*}))\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*},
\nonumber\\ \label{eq_marginal_1}
\end{eqnarray}
so we can then safely identify
\begin{eqnarray}
P(det|\lambda) & \equiv & f(I(\alpha,\alpha^{*})),
\end{eqnarray}
with $f(I(\alpha,\alpha^{*})) \geq 0$, $\forall I(\alpha,\alpha^{*})$.
This last is nothing but solving a linear system with only one restriction and an
infinite number of free parameters, whose subspace of solutions intersects the region
$f(I(\alpha,\alpha^{*})) \geq 0$ $\forall \alpha,\alpha^{*}$: see Appendix \ref{Math}.
Of course, once at this point, joint detection probabilities would in principle
impose additional restrictions on $f(x)$; for detectors $i,j$ we would have,
for some constant $K_{(j)}$ (``joint''),
\begin{eqnarray}
&& K_{(j)} \int
(I_i(\alpha,\alpha^{*}) - I_{0,i})\cdot(I_j(\alpha,\alpha^{*}) - I_{0,j})
\nonumber\\
&&\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad
\times \ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*} \nonumber\\
&&\quad
=
\int
f(I_i(\alpha,\alpha^{*}))\cdot f(I_j(\alpha,\alpha^{*}))
\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*}.
\nonumber\\ \label{eq_joint_1}
\end{eqnarray}
\hfill\endproof
\vspace{0.2cm}\noindent
(B)
In principle, it is unclear whether all possible $f(x)$'s satisfying (\ref{eq_marginal_1})
can be made to meet (\ref{eq_joint_1}), therefore we propose:
instead of demanding (\ref{eq_joint_1}) from the beginning, we define a new
function $\Gamma(x,y)$, so that
\begin{eqnarray}
&& K_{(j)} \int
(I_i(\alpha,\alpha^{*}) - I_{0,i})\cdot(I_j(\alpha,\alpha^{*}) - I_{0,j})
\nonumber\\
&&\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad
\times \ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*} \nonumber\\
&&\quad\quad
=
\int
\Gamma(I_i(\alpha,\alpha^{*}),I_j(\alpha,\alpha^{*}))
\ W(\alpha,\alpha^{*})\ d\alpha d\alpha^{*}, \nonumber\\
\label{eq_joint_2}
\end{eqnarray}
so we can now identify
\begin{eqnarray}
P_{i,j}(det|\lambda)
&\equiv& \Gamma(I_i(\alpha,\alpha^{*}),I_j(\alpha,\alpha^{*})).
\end{eqnarray}
In Appendix \ref{Math} we have shown that it is always possible to find some
suitable $0 \leq \Gamma(x,y) \leq 1$, for all $x = I_i(\alpha,\alpha^{*})$ and
$y = I_j(\alpha,\alpha^{*})$, i.e., for all possible pairs $\alpha,\alpha^{*}$.
The absence of factorability on the $\alpha$'s, i.e., the fact that (we may have)
\begin{eqnarray}
&&\Gamma(I_i(\alpha,\alpha^{*}),I_j(\alpha,\alpha^{*})) \neq \nonumber\\
&&\quad\quad\quad\quad\quad\quad
f(I_i(\alpha,\alpha^{*}))\cdot f(I_j(\alpha,\alpha^{*})),
\end{eqnarray}
may come as a surprise to some, given than Clauser-Horne factorability \cite{CH74}
on the hidden variable $\lambda$ is usually taken for granted; this is a mistake
\cite{factorability}.
We have tried to clarify the whole issue in Appendix \ref{Rev}:
in general, we have shown that that non-factorability, aside from perfectly
legitimate from the mathematical perspective, is in this case possibly related
with the need to include new hidden variables in the model.
\hfill\endproof
\vspace{0.2cm}
It is now convenient, for the following sections, to make our notation lighter,
with $\alpha \equiv \alpha,\alpha^{*}$, and also to define
\begin{eqnarray}
\hat{f}_i( \alpha ) &\equiv& f( I_{i}(\alpha)), \label{f_hat}\\
\hat{\Gamma}_{i,j}( \alpha ) &\equiv& \Gamma(I_i(\alpha),I_j(\alpha)). \label{Gamma_hat}
\end{eqnarray}
\section{\label{Pol}
Adding intermediate devices:
polarizers, PBS's...}
Once we place one or more devices between the crystal and the detectors,
typically polarizers, polarizing beam splitters (PBS) or other devices to allow
polarization measurements (such as in Fig. \ref{Scheme}),
in general we cannot any longer describe the fields between both with only one
set $\{\alpha\}$ of mode-amplitudes;
we need to redefine our $\alpha$'s as now associated to a particular position
$\mathbf{r}$ (they do not any longer determine a frequency mode for all
space \cite{note_modes}).
Hence, we will now have
\begin{eqnarray}
\alpha(\mathbf{r}) \equiv
\{ \alpha_{\mathbf{k},\gamma}(\mathbf{r}), \alpha_{\mathbf{k},\gamma}^{*}(\mathbf{r}) \},
\end{eqnarray}
and, letting $\mathbf{r}_s$ be the position of the source (the crystal), and
$\mathbf{r}_i$ the position of the $i$-th polarizer or PBS (or any other
intermediate device), we will also redefine
\begin{eqnarray}
\alpha_s &\equiv& \alpha(\mathbf{r}_s), \quad
\alpha_i \equiv \alpha(\mathbf{r}_i),
\end{eqnarray}
with $\alpha_i$ including the \emph{relevant} amplitudes at the (empty) exit
channels of that $i$-th intermediate device.
With $\alpha_s,\alpha_i$ corresponding each to (a set of) modes with different
(sets of relevant) wavevectors $\{\mathbf{k}_s\}$ and $\{\mathbf{k}_i\}$ \cite{note_wvectors},
we can then regard them as two (sets of) statistically independent random variables,
with all generality.
The intensity at the entrance of detector $i$-th will therefore depend now not only on
$\alpha_s$ but also on $\alpha_i$:
\begin{eqnarray}
P_{i}
&\propto& \langle I_i(\alpha_s,\alpha_i) - I_{0,i} \rangle \nonumber\\
&=& \int_{\alpha_s} \int_{\alpha_i}
( I_{i}(\alpha_s,\alpha_i) - I_{0,i} )\ W(\alpha_s) \ W(\alpha_i)\ d\alpha_s d\alpha_i.
\nonumber\\ \label{marginal_p}
\end{eqnarray}
Once more, something like that can always be rewritten (see former section),
for some suitable and positively defined $f^{\prime}(x)$, as
\begin{eqnarray}
P_{i}
&=& \int_{\alpha_s} \int_{\alpha_i}
f^{\prime}( I_{i}(\alpha_s,\alpha_i))\ W(\alpha_s) \ W(\alpha_i)\ d\alpha_s d\alpha_i.
\nonumber\\
\end{eqnarray}
Integrating on $\alpha_i$ we would obtain
\begin{eqnarray}
P_{i} &=& \int_{\alpha_s} \hat{f}^{\prime}_{i}( \alpha_s )\ W(\alpha_s) \ d\alpha_s,
\label{f_prime_hat}
\end{eqnarray}
from where we define a new function $\hat{f}^{\prime}_{i}( \alpha_s )$.
On the other hand, for joint detections we would have
\begin{eqnarray}
P_{i,j} &\propto&
\int_{\alpha_s} \int_{\alpha_i} \int_{\alpha_j}
( I_{i}(\alpha_s,\alpha_i) - I_{0,i} ) \cdot ( I_{j}(\alpha_s,\alpha_j) - I_{0,j} )
\nonumber\\ &&\quad\quad\quad\quad\quad
\times\ W(\alpha_s) \ W(\alpha_i)\ W(\alpha_j)\ d\alpha_s d\alpha_i d\alpha_j,
\nonumber\\ \label{joint_p}
\end{eqnarray}
which again can always be rewritten (again see former section), for
some positively defined $\Gamma^{\prime}(x,y)$, as
\begin{eqnarray}
P_{i,j} &=& \int_{\alpha_s} \int_{\alpha_i} \int_{\alpha_j}
\Gamma^{\prime}( I_{i}(\alpha_s,\alpha_i), I_{j}(\alpha_s,\alpha_j)) \nonumber\\
&&\quad\quad\quad\quad\quad
\times\ W(\alpha_s) \ W(\alpha_i)\ W(\alpha_j)\ d\alpha_s d\alpha_i d\alpha_j,
\nonumber\\
\end{eqnarray}
and integrating on $\alpha_i,\alpha_j$ we would obtain
\begin{eqnarray}
P_{i,j} &=& \int_{\alpha_s} \hat{\Gamma}^{\prime}_{i,j}( \alpha_s )\ W(\alpha_s) \ d\alpha_s,
\label{Gamma_prime_hat}
\end{eqnarray}
from where we can again define yet another new probability density
function $\hat{\Gamma}^{\prime}_{i,j}( \alpha_s )$.
It is interesting for the sake of clarity to compare the two situations: with (primed
functions) and without polarizers (unprimed).
It is easy to see that, because the detector only sees the intensity at its entrance
channel, clearly (we drop the ``s'' subscript for simplicity),
\begin{eqnarray}
f^{\prime}(x) &=& f(x), \quad \forall x, \\
\Gamma^{\prime}(x,y) &=& \Gamma(x,y), \ \forall x,y.
\end{eqnarray}
while, naturally, in general
$\hat{f}^{\prime}_{i}( \alpha ) \neq \hat{f}_{i}( \alpha )$,
as well as
$\hat{\Gamma}^{\prime}_{i,j}( \alpha ) \neq \hat{\Gamma}_{i,j}( \alpha )$.
\section{\label{Fac}On non-factorability}
\subsection{\label{Math_Fac}Mathematical analysis}
Let us go back to the case with just the source and the detectors; we will soon
see the following does nevertheless also apply when polarizers or other devices
are added to the setup, just the same.
According to our reasonings in App. \ref{Rev}, and using (\ref{f_hat})--(\ref{Gamma_hat}),
we now realize that there is no way to avoid
\begin{eqnarray}
\hat{\Gamma}_{i,j}(\alpha) = \hat{f}_{i}(\alpha) \cdot \hat{f}_{j}(\alpha),
\end{eqnarray}
unless we introduce some additional dependence of the kind
$\hat{f}(\alpha) \rightarrow \hat{\hat{f}}(\alpha,\mu)$,
so that then
\begin{eqnarray}
\hat{\hat{\Gamma}}_{i,j}(\alpha,\mu)
\neq \hat{\hat{f}}_{i}(\alpha,\mu) \cdot \hat{\hat{f}}_{j}(\alpha,\mu),
\end{eqnarray}
where we add a second ``hat'' to avoid an abuse of notation, and where $\mu$
stands for a new set of random variables.
This $\hat{\hat{f}}(\alpha,\mu)$ should be interpreted as a detection
probability conditioned to the new vector of random variables $\mu$, i.e,
\begin{eqnarray}
\hat{\hat{f}}(\alpha,\mu) \equiv P(det|\alpha,\mu).
\end{eqnarray}
We will impose further demands on $\hat{\hat{f}}(\alpha,\mu)$, defining
\begin{eqnarray}
\hat{\hat{f}}(\alpha,\mu) \equiv f(I(\alpha,\mu)),
\end{eqnarray}
something forced by strictly physical arguments: the choice $f(I(\alpha,\mu))$ must
prevail over other possible ones - for instance $f(I(\alpha),\mu)$ - due to the need
to respect the dependence of the probabilities of detection (conditioned to $\alpha$
or not) alone on the intensity that arrives to the detector, and nothing else.
Now, with the density function $\rho_{\mu}(\mu)$, we could write
\begin{eqnarray}
P(det|\alpha) &=& \int_{\mu} P(det|\alpha,\mu)\ \rho_{\mu}(\mu)\ d \mu, \\
P_{i,j}(det|\alpha) &=& \int_{\mu} P_{i,j}(det|\alpha,\mu)\ \rho_{\mu}(\mu)\ d \mu ,
\end{eqnarray}
allowing us to recover our former definitions (\ref{f_hat})--(\ref{Gamma_hat}):
\begin{eqnarray}
\hat{f}(\alpha) \equiv P(det|\alpha), \\
\hat{\Gamma}_{i,j}(\alpha) \equiv P_{i,j}(det|\alpha).
\end{eqnarray}
For joint detections, the additional variable $\mu$ is particularly relevant
because, we will always have that while
\begin{eqnarray}
P_{i,j}(det|\alpha,\mu) = P_{i}(det|\alpha,\mu) \cdot P_{j}(det|\alpha,\mu),
\end{eqnarray}
in general
\begin{eqnarray}
P_{i,j}(det|\alpha) \neq P_{i}(det|\alpha) \cdot P_{j}(det|\alpha),
\end{eqnarray}
or we could equivalently say that while necessarily
\begin{eqnarray}
\hat{\hat{\Gamma}}_{i,j}(\alpha,\mu) = \hat{\hat{f}}_{i}(\alpha,\mu) \cdot \hat{\hat{f}}_{j}(\alpha,\mu),
\end{eqnarray}
in general
\begin{eqnarray}
\hat{\Gamma}_{i,j}(\alpha) \neq \hat{f}_{i}(\alpha) \cdot \hat{f}_{j}(\alpha),
\end{eqnarray}
where of course
\begin{eqnarray}
\hat{\Gamma}_{i,j}(\alpha) = \int_{\mu}
\hat{\hat{f}}_{i}(\alpha,\mu) \cdot \hat{\hat{f}}_{j}(\alpha,\mu)\ \rho_{\mu}(\mu)\ d\mu.
\label{int_mu}
\end{eqnarray}
To conclude this section, we recover the case with intermediate devices:
due to $\alpha_i,\alpha_j$ being, as defined, independent from one another and also
from $\alpha_s$, our hypothetical ``flag'' $\mu$ cannot be associated with none of them.
Therefore, in general, and in principle, not only
\begin{eqnarray}
\hat{\Gamma}^{\prime}_{i,j}( \alpha ) \neq
\hat{f}^{\prime}_{i}( \alpha ) \cdot \hat{f}^{\prime}_{j}( \alpha ),
\label{Gamma_prime_nofac}
\end{eqnarray}
but also
$\hat{\Gamma}_{i,j}( \alpha ) \neq \hat{f}_{i}( \alpha ) \cdot \hat{f}_{j}( \alpha )$
either.
We use ``in principle'' because this question is not yet analyzed
in detail;
we now see crystal clear, though, that this possible non-factorability
on $\alpha$'s is nothing more than an internal feature of the model's
mathematical structure,
bearing no relevance in regard to its double-sided compatibility (or
absence of it) both with local-realism and the quantum predicted
correlations.
Indeed, the crucial element in that sense is the intervention of new
random vacuum amplitudes introduced by the new devices in the setup,
as we will see in Sec. \ref{Comp}.
\subsection{\label{Phys_int}A physical interpretation}
Recapitulating, we have that while
$P_{i,j}(det|\alpha,\mu) = P_{i}(det|\alpha,\mu) \cdot P_{j}(det|\alpha,\mu)$,
however $P_{i,j}(det|\alpha) \neq P_{i}(det|\alpha) \cdot P_{j}(det|\alpha)$,
where $\mu$ would be a new, additional HV (or vector of HV's) and should, in
the ultimate term, correspond to some physically relevant variable that we
have so far overlooked, perhaps containing additional information about the state
of the crystal and the laser.
Indeed, if it was eventually necessary to introduce a new variable $\mu$,
physical intuition suggests that it could probably be associated to the
phase of the laser complex amplitude \cite{laser_phase};
with the laser in a coherent state, it would be justified to approximate
$\rho_{\mu}(\mu) \approx \delta(\mu_0)$, where $\mu_0$ would be the phase or the
quasi-classical wave.
Perhaps it is needless to add that in such a case expressions such as (\ref{int_mu})
or a similar one for $\Gamma^{\prime}$'s would still be fully operative, as long
as the dependence on $\mu_0$ is ``absorbed'' in the functional dependence at the
left side of the equation.
Anyway, whether this $\mu$ has an obvious physical meaning or not should not
bother us too much here, as it does not affect at all the correctness and/or
the implications of our mathematical results.
\section{\label{Comp}Complementary questions}
\subsection{\label{LR_v_QM}
Wigner-PDC's local realism vs. quantum correlations}
Though former mathematical developments are fully meaningful and self-contained on
their own, yet it would be convenient to give some hints on how a local-realist (LR)
model can account for typically quantum correlations, which are known to defy that
very same local realism (LR).
In the first place and as a general answer,
what the Wigner-PDC picture proves is that LR is respected by a certain subset of
all the possible quantum states, specifically the ones that can be generated from
a non-linear mix of the QED-vacuum (which therefore acts as an ``input'' for the model)
with a quasi-classical (a high-intensity coherent state), highly directional signal,
the laser ``pump'' (which indeed enters in the model as a non-quantized, external
potential \cite{pdc1}).
Moreover, such a restriction is clearly not arbitrary at all, since it arises from
a very simple quantum electrodynamical model of the process of generation of
polarization-entangled pairs of photons from Parametric Down Conversion (PDC):
see for instance eq. (4.2) in \cite{pdc1}.
\subsection{\label{Rates}
Detection rates and ``efficiencies''}
Aside from subscripts, we will now also drop ``hats'' and ``primes'' for simplicity;
of course the fact that $\Gamma_{i,j}(\alpha)$ may not be in general $\alpha$-factorisable,
\begin{eqnarray}
\Gamma_{i,j}(\alpha) \neq f_i(\alpha) \cdot f_j(\alpha),
\end{eqnarray}
does not at all mean that it cannot well satisfy
\begin{eqnarray}
&&\int \Gamma_{i,j}(\alpha) \ W(\alpha)\ d\alpha = \nonumber\\
&&\quad\ \
\left[ \int f_i(\alpha)\ W(\alpha)\ d\alpha \right] \cdot
\left[ \int f_j(\alpha^{\prime})\ W(\alpha^{\prime})\ d\alpha^{\prime}
\right], \nonumber\\
\end{eqnarray}
i.e. (let us from now use superscripts ``W'' and ``exp'' to denote, respectively,
theoretical and experimental detection rates):
\begin{eqnarray}
P_{i,j}^{(W)}(det) = P_{i}^{(W)}(det) \cdot P_{j}^{(W)}(det),
\end{eqnarray}
which is indeed the sense in which the hypothesis of ``error independence''
is introduced, to our knowledge, in every work on LHV models \cite{LHVs}.
This sort of conditions over ``average'' probabilities (average in the sense
that they are integrated in the hidden variable, may that be $\alpha$ alone
or also some other one) are the only ones that can be tested in the actual experiment;
there, we can just rely on the number of counts registered on a certain
time-window $\Delta T$, and the corresponding estimates of the type
\begin{eqnarray}
P_{i}^{(exp)}(det) \approx \frac
{ n.\ joint \ det.\ (i,j) \ in\ \Delta T }
{ n.\ marg. \ det.\ (j) \ in\ \Delta T}. \label{estimation_marg}
\end{eqnarray}
Now, if we wish to include some additional uncertainty element reflecting the
technological limitations (a ``detection efficiency'' parameter), what we have
to do is to redefine the overall detection probabilities as
\begin{eqnarray}
P_{i}^{(exp)}(det) &\equiv& \hat{\eta}_i \cdot P_{i}^{(W)}(det), \label{eta_marginal}\\
P_{i,j}^{(exp)}(det) &\equiv& \hat{\eta}_i\hat{\eta}_j \cdot P_{i,j}^{(W)}(det), \label{eta_joint}
\end{eqnarray}
as well as
\begin{eqnarray}
P_{i}^{(exp)}(det|\alpha) &\equiv& \hat{\eta}_i \cdot P_{i}^{(W)}(det|\alpha),
\label{eta_red_m} \\
P_{i,j}^{(exp)}(det|\alpha) &\equiv& \hat{\eta }_i\hat{\eta}_j \cdot P_{i,j}^{(W)}(det|\alpha),
\label{eta_red_j}
\end{eqnarray}
where $0 \leq \hat{\eta}_i \leq 1$ would play the role of such an (alleged) detection
efficiency, the ``hats'' remarking the fact that the customary definition of the
analogous quantity in QInf involves not only our $\hat{\eta}$'s but also the
non-technological contribution.
From the point of view of the experimenter it is very difficult to isolate both
components (Glauber's theory \cite{detection} does not predict a unit detection
probability even for high intensity signals);
we should perhaps then confine ourselves to the term ``observable detection rate''
instead of using the clearly misleading one of ``detector inefficiency''.
\subsection{
Consequences on Bell tests, their supplementary assumptions and critical efficiencies}
That said and going to a lowest level of detail, states in such an (LR) subset
of QED can still indeed exhibit correlations of the class that is believed to collide
with LR, yet the procedure through which they are extracted from the experimental set
of data does not meet one of the basic assumptions required by every test of a Bell
inequality:
they do not keep statistical significance with respect to the physical set of
``states'' or hidden instructions \cite{LHVs}).
To guarantee that statistical significance we must introduce some of the following
two hypothesis:
\vspace{0.2cm}\noindent
(i) all coincidence detection probabilities are independent of the polarizers'
orientations $\phi_i,\phi_j$ (this is what we call ``fair-sampling'' \cite{fs},
for a test of an homogeneous inequality \cite{hom}), which implies
\begin{eqnarray}
P_{ij}(det|\phi_i, \phi_j, \alpha) = P_{ij}(det|\alpha_s),
\end{eqnarray}
where of course (see Sec. \ref{Pol}) $\alpha \equiv \alpha_s \oplus \alpha_i \oplus \alpha_j$,
and where we recall that $\phi_i,\phi_j$ would determine which vacuum modes amongst the
sets $\alpha_i,\alpha_j$ would intervene in the detection process,\\
\vspace{0.2cm}\noindent
(ii) the interposition of an element between the source and the detector cannot
in any case enhance the probability of detection (the ``no-enhancement'' hypothesis
\cite{no_enhancement}, needed to test the Clauser-Horne inequality \cite{CH74}, and
presumably every other inhomogeneous one \cite{note_any}),
\begin{eqnarray}
P_{i}(det|\phi_i, \alpha) \leq P_{i}(det|\infty, \alpha_s).
\end{eqnarray}
with $\infty$ denoting the absence of polarizer and with
$\alpha \equiv \alpha_s \oplus \alpha_i$ this time.
\vspace{0.2cm}\noindent
Following our developments in Sec. \ref{Pol} one can easily see that (i) is not
in general true, and according to \cite{WignerPDC_enhancement} neither is (ii).
I.e., whenever states of light are prepared so as to produce the sort of quantum
correlations that are known to defy LR, these last come supplemented with the necessary
features that prevent it from happening... how could it not?
Yet, a mere breach of (i) or (ii) is not enough to assert the existence of a Local Hidden
Variables (LHV) model, which is an equivalent way of saying that the results of the
experiment respect LR: it is more than well known that this can only happen for certain
values of the observed detection rates \cite{LHVs,crit_eff}.
However,
from the point of view of (\ref{eta_red_m})--(\ref{eta_red_j}), and given the fact
that, as proven in Secs \ref{No_pol} and \ref{Pol}, the Wigner-PDC
is in all circumstances in accordance with expressions (\ref{A_loc_causal})--(\ref{B_loc_causal}),
and hence to all possible Bell inequalities whether our $\hat{\eta}$'s are
equal or less than unity,
the so-called ``critical efficiencies'' would merely stand, at least as far as
PDC-generated photons are concerned, for bounds on the detection rates that we can
experimentally observe (these last in turn constrained by the only subset of quantum
states that we can physically prepare).
\subsection{\label{Kot}A recent related test}
The recent test in \cite{Kot_et_al12} would seem formally equivalent to a Bell
test of an homogeneous inequality \cite{hom}, but for two questions:
(i) they do not require remote observers;
(ii) the use of analogical measurements, that would seem to exclude the
so-called ``detection loophole'' that is widely recognized in other tests.
The purpose of this section is to show that such loophole still remains,
though making it evident requires a subtler approach, according to our approach
here one perhaps more realistic than the usual assumption of ``random errors''.
Besides, and to my knowledge, \cite{Kot_et_al12} is the only amongst other
recent related tests \cite{Peruzzo12, Kaiser12} which does not acknowledge,
explicitly, one or other loophole leading to compatibility with the
local-realistic interpretation.
Basically,
Kot {\it et al}'s proposal in \cite{Kot_et_al12} rests on the probing of some
``test function'' $F \equiv F(Q_{1},Q_{2},\ldots Q_{N})$,
where $Q_{m}$ is an outcome obtained when an observable $\hat{Q}_{m}$ is
measured, and where $\{\hat{Q}_{m}\}$ is a set of mutually exclusive
observables.
Later, they will be concerned with an ``average'' $\beta = \langle F \rangle$,
which, if a local hidden variables (LHV) model turned out to exist, would
have to be expressible as
\begin{eqnarray}
\langle F \rangle_{\Lambda} =
\int_{\Lambda}
F(\{Q_m\}) \cdot P( \{Q_m\} | \lambda ) \ \rho(\lambda) \ d\lambda,
\label{eq_0}
\end{eqnarray}
with $\lambda$ as usual a vector of hidden variables and $\Lambda$ the overall
space of events.
The quantity $P( \{Q_m\} | \lambda ) = f( Q_{1},\ldots Q_{N} | \lambda )$
would stand for a joint probability density function (conditioned to the
state $\lambda$) for any set of (predetermined) results for any set of possible
measurements upon the observables $\{\hat{Q}_m\}$.
Each $\hat{Q}_{m}$ will be then sampled on a different subset
$\Lambda_m \subset \Lambda$... which would be no problem as long as
all $\{\Lambda_m\}$ are statistically faithful to $\Lambda$;
however, the fact that measurements may not always give a result (or that
these measurements are effectively completed at different time-stamps)
may destroy that statistical significance, hence invalidating that of the
estimate of $\beta$.
To show that this may happen even in view of (ii), we must go to Glauber's
expression in (\ref{Glauber_s}).
The key is that, for a given time-stamp, in general $P(t) < 1$, though of
course after a certain interval $\Delta t$ the accumulated probability,
\begin{eqnarray}
P_{det}(\Delta t) = \int_{t}^{t + \Delta t} P_{det}(\tau) d\tau,
\label{p_acc}
\end{eqnarray}
may approach unity, something irrelevant for our argument:
the hypothetical $0$-instructions in a hypothetical LHV model (see for instance
\cite{LHVs}) would no longer have anything to do with some ``detection inefficiency'',
but simply express the fact that for some given time-stamp and observable,
$P(t)$ may be less than unity.
Other models of ``detection'' may involve a time-integral
of the intensity at the entrance of the detector, or other sophistications;
they would not affect our argument as long as we recognize (\ref{Glauber_s}) as
a time-dependent quantity, in general satisfying $P(\mathbf{r},t) < 1$.
Once here, the physical connection with the test in \cite{Kot_et_al12} can be
established by assigning to each time stamp $t$ a set ${\cal M}_{t}$ of hidden
instructions of the usual kind (i.e., a different LHV model ${\cal M}_{t}$
for each $t$): again see \cite{LHVs}.
In particular, following \cite{prev07}, a detection on the ``signal'' arm, at a
time-stamp $t$, prompts the analysis of the signal (coming from the homodyne setup
and entering a high frequency oscilloscope) at the ``idler'' one, over a fixed
time window.
Yet, a correlation between the detection time-stamp at the
signal arm and the choice of observable at the idler would seem to require
communication or ``signaling'' between the two measurement setups.
This last difficulty, which would burden our argument with the need for an
additional ``locality loophole'', can, however, be overcome too:
consider an observable $\hat{Q}$ that (always) produces two possible results
$Q = \{q_1,q_2\}$.
For simplicity let us this time denote $P_{Q}(q|\lambda)$ the probability of
an outcome $q$ when $Q$ is measured on the state $\lambda$.
Then,
\begin{eqnarray}
\langle Q \rangle_{\Lambda} = \frac{
q_1 \cdot P_{Q}(q_1|\Lambda) + q_2 \cdot P_{Q}(q_2|\Lambda)}
{P_{Q}(det|\Lambda)};
\end{eqnarray}
however, if each result is associated to a detection at a different set of
time-stamps, $\{t_{i,1}\},\{t_{i,2}\}$, then the experimentally accessible
quantity is
\begin{eqnarray}
\langle Q \rangle_{ob} = \frac{
q_1 \cdot P_{Q}(q_1|\Lambda^{(Q)}_1) + q_2 \cdot P_{Q}(q_2|\Lambda^{(Q)}_2)}
{P_{Q}(det|\Lambda)}.
\end{eqnarray}
In this last expression,
$\Lambda^{(Q)}_1,\Lambda^{(Q)}_2 \subset \Lambda$ are again two
different sets selected by the correlation between the time-stamp at
one arm and the result of the measurement at the other when $\hat{Q}$
is measured.
According to the approach in \cite{pdc1,pdc2,pdc3,pdc4,pdc5,pdc6,pdc7},
the set of relevant vacuum electromagnetic modes inserted at the source are
still contained in the fields arriving at each detector, and this will
indeed induce some correlation between detection time-stamp at one side
and a measurement's outcome at the other, and in general between any two
events which are related to the intensity of the incoming signal.
This induced correlation is also at the core of other recent results
in \cite{pdc8,pdc9}.
A more detailed argument can be obtained by e-mail from the author.
\subsection{\label{Sub}
An approach to realistic detectors and average ZPF subtraction}
This section approaches some physical considerations with the aim
of showing that there is plenty of room for a suitable physical interpretation of the
model, even when that is not strictly necessary for the coherence of our results,
at least from the purely mathematical point of view.
Indeed, we have already descended to the physical level when we established,
in former sections, the dependence of our $f$'s and $\Gamma$'s solely on the
intensity arriving at the detector.
Expectable behavior for a physical device would typically include a ``dead-zone'',
an approximately linear range and a ``saturation'' at high intensities (this is
indeed the kind of behavior suggested for instance in \cite{W_LHV_02});
amongst other restrictions this would imply, for instance, $f(I(\alpha)) = 0$, when
$I(\alpha) \leq \bar{I}_0$ (this last a threshold that may even surpass the expectation
value of the ZPF intensity),
as well as $\Gamma(I_i(\alpha),I_j(\alpha)) = 0$ either for $I_i(\alpha)$ or $I_j(\alpha)$
below $\bar{I}_0$.
Neither these restrictions nor other similar ones would in principle invalidate our
proofs in Appendix \ref{Math}, which seem to provide room enough to simulate a wide
range of possible behaviors;
however, we must remark that none of our $f$'s and $\Gamma$'s can ultimately be
considered as fully physical models, due to the fact that they represent point-like
detectors
(the implications of such an over-simplification may become clearer in the light
of our following point here).
In close relation to the former, we also propose here a simple physical interpretation
of the term $-I_0$ appearing in the expressions for the detection probabilities.
From the mathematical point of view such subtraction arises from a mere manipulation of
Glauber's original expression \cite{detection}.
From the physical one, a realistic interpretation would be more than desirable, as that
subtraction of ZPF intensity is for instance crucial to explain the absence of an
observable contribution of the vacuum field on the detectors' rates \cite{non_detec_ZPF};
of course we mean ``explain in physical terms''; from the mathematical point of view
our model here already predicts a vanishing detection probability for the ZPF alone.
My suggestion is that the $-I_0$ term must be (at least) related to the average flux
of energy going through the surface of the detector in the opposite direction to the
signal (therefore \emph{leaving} the detector).
That interpretation fits the picture of a detector as a physical system producing
a signal that depends (with more or less proportionality on some range) on the
total energy (intensity times surface times time) that it accumulates.
I am just saying ``at least related'', bearing in mind that to establish
such association we would first have to refine the point-like model of a detector
which stems directly from the original Glauber's expression \cite{detection}.
\section{\label{Conc}Conclusions}
We have shown that the Wigner description of PDC-generated (Parametric Down Conversion)
photon-entanglement, so enthusiastically developed in the late nineties
\cite{pdc1,pdc2,pdc3,pdc4,pdc5,pdc6,pdc7} but then ignored in recent years,
can be formulated as entirely local-realistic (LR).
A formalism that is one-to-one with a quantum (field-theoretical) model of the
experimental setup can be cast, thanks to an additional manipulation (also one-to-one),
into a form that respects all axiomatic laws of probability \cite{p_ax}, and therefore
LR, as defined for instance in Sec. \ref{Loc}.
The original quantum electrodynamical model takes as an input the vacuum state,
which accepts a well defined probabilistic description through the so-called Wigner
transform: this is the fact that the analogy with a local-realistic theory
is conditioned to.
What we call Wigner-PDC accounts, then, for a certain subset within the space of all
possible QED-states, determined by a particular set of initial conditions and a certain
Hamiltonian governing the time-evolution \cite{c_Ham},
restrictions that seem to guarantee (according to us here) the compatibility with LR.
Aiming for the maximum generality, as well as to avoid some possible (still under
examination) difficulties with factorisable expressions, we have renounced to what we
call $\alpha$-\emph{factorability} of the joint detection expressions.
Such a choice is not only perfectly legitimate \cite{factorability}, but may also be
supported by a well feasible interpretation (see Appendix \ref{Fac});
nevertheless, further implications of that non-factorability on $\alpha$ will also be
left to be examined elsewhere.
Again, whatever they finally turn out to be, they are also irrelevant for our main
result in this paper:
the Wigner-PDC formalism can be cast into a form that respects all axiomatic laws of
probability for space-like separated events.
Neither does the explicit distinction between the cases where polarizers (or other devices
placed between the source and the detectors) are or not included in the setup introduce
any conceptual difference from the point of view of our main result;
however, the question opens room to remark some of the main differences of the Wigner-PDC
model (actually, also its Hilbert-space analogue) with the customary description used
in the field of Quantum Information (QInf):
here, each new device introduces noise, new vacuum amplitudes that fill each of its
empty polarization channels at each of its exists, in contrast to the usual QInf ``black
box'', able to extract polarization information from a photon without any indeterministic
component.
As a matter of fact, those additional random components hold the key to explain the
variability of the detection probability (see note \cite{WignerPDC_enhancement}) that
is necessary, from the point of view of Bell inequalities, to reconcile quantum
predictions and LR (see Sec. \ref{LR_v_QM}).
In particular, the immediacy with which the phenomenon of ``detection probability
enhancement'' arises in the Wigner-PDC framework would suggest that this may be after
all the right track to understand why after several decades the minimum
detection rates (or, in QInf terminology, critical detection efficiencies) that
would lead to obtain conclusive evidence of non-locality are yet to be reached
\cite{crit_eff}.
In Sec.\ref{Kot} we have also addressed a recent no-go test that I consider particularly
relevant, as it does not explicitly acknowledge any loophole, and which could also be
given a local-realistic interpretation in the light of the present framework.
In Sec. \ref{Sub} we have done a first, general approach on the question of whether
our reinterpretation of the detection probabilities is consistent with the actual
physical behavior of detectors.
A closer look to this question is nevertheless left out of the scope of the paper;
former proposals \cite{W_LHV_02} in regard to this issue aimed perhaps too
straightforwardly to the physical level, while they did not even guarantee consistency
with the framework we have settled here
(proof of this is that it introduces divergences in relation to the purely quantum
predictions, divergences later experimentally disproved for instance in \cite{Brida_et_al02}).
In close relation with the former, we have also suggested a possible simple
physical interpretation of $I_0$-subtraction taking place in the expressions
for the detection probabilities:
work in any of this two directions would anyway require a departure from the point-like
model of a detector on which we have focused here.
Summarizing, we have shown that a whole family of detection models
\begin{eqnarray}
{\cal M}_{det} \equiv \{f(I(\alpha,\mu)),\Gamma(I_i(\alpha,\mu),I_j(\alpha,\mu)\},
\end{eqnarray}
can be found, consistent both with the quantum mechanical expressions from the
Wigner-PDC model and LR.
A close examination of the constraints coming from the physical behavior of the detectors
and other experimentally testable features is left as a necessary step for the future,
with the aim of establishing a subset
physically feasible ones;
nevertheless, we have the guarantee that all of them produce suitable predictions,
as so does their quantum electrodynamical counterpart.
Yet, even at the purely theoretical level some other features remain open
too: as a fundamental one, to what extent the model
requires what we have called non-factorability on $\alpha$'s.
Once more and after all,
QM is just a theory, a theory that provides a formalism upon which to build
models, models than can (and should) be refined based on experimental evidence,
something that (again) also applies just as well to the one we are dealing with here
\cite{note_WPDC}.
Perhaps the particle properties of light are not enough to assume that the
current model of ``a photon'' is the best representation (and most complete)
that we can achieve of light;
many of those properties can be understood, I believe, from entirely classical models
\cite{classic_photon}, as well.
Quantum entanglement seems to manifest in many of its ``reasonable'' features but
at least as (this particular model of) Parametric Down Conversion is concerned
and so far, whenever local-realism would seem to be challenged new phenomena can
be invoked so as to prevent, at least potentially, that possibility.
Such are ``unfair sampling'', ``enhancement'' (as a particular case of variable
detection probability) and over all detection rates low enough to open room for
the former two.
Those phenomena find theoretical support in the Wigner-PDC picture, but
definitely not in the usual, based merely on the correspondence
between the $\tfrac{1}{2}$-spin algebra for massive particles and the
polarization states of a plane wave
(a photon then looks just as a $\tfrac{1}{2}$-spin particle, what some like to
call a two-level system, but for the magnitude of the angular momentum it
carries, and its statistics of course).
To explore, and exhaust if that is the case, alternative routes such as the one
here is not only sensible but also necessary.
\section*{Acknowledgments}
I thank R. Risco-Delgado, J. Mart{\'i}nez and A. Casado for very useful discussions,
though most of them at a much earlier stage of the manuscript.
They may or may not agree with either the whole or some part of my conclusions and
results, of course.
|
\section{Introduction}
\label{intro}
Solar filaments (prominences) are cold dense plasmas suspended in the corona. Their fine threads are
seen in emission as bright prominences at the limb and in absorption as dark filaments against the disk.
A filament is formed and maintained above the magnetic polarity inversion line, in a magnetic
structure called a filament channel, in which the filament can be supported by the magnetic field.
For filament formation and maintenance, \citet{mart98} gave a comprehensive review,
including the filament structure, chirality, magnetic topology, and mass flows. Filaments would erupt
at the onset of MHD instabilities or loss of equilibrium. It is now widely accepted that filament eruption
is often associated with other solar activities, such as solar flares and coronal mass ejections (CMEs).
\citet{martens89} considered filaments as the coronal part of an electric
current that loses MHD equilibrium at the flare onset and starts to erupt outwards. This leads directly
to the observed CME. \citet{kaas85} showed that as the filament moves upwards, a neutral line is
formed beneath it, which becomes the site of magnetic reconnection and particle acceleration
during the flare. The accelerated particles precipitate into the chromosphere along the field lines,
and produce hard X-ray, UV, and optical (such as H$\alpha$) emission forming flare ribbons. The energy deposition
in the lower atmosphere further causes chromospheric evaporation, which leads to the observed coronal extreme-ultraviolet
(EUV) and soft X-ray emissions \citep{canf82}.
It has been proposed that two filaments (or filament segments) of the same chirality, when approaching each other,
can merge to form long filaments or interact with each other \citep{daza48, martens01}. This has been confirmed by
observations \citep{schm04} and numerical simulations \citep{devo05, aula06}. However, in these observations
and numerical experiments, the authors found no initiation of an eruption. Theoretically,
\citet{martens01} presented a ``head-to-tail'' linkage model for the eruption of filaments. Furthermore,
\citet{ural02} suggested that the dual-filament interaction could cause solar eruptions. It is
fortunate that there are indeed a few observations indicating that
the linkage of two filaments can initiate an eruption. \citet{sujt07} presented new observations
of the interaction of two nearby but distinct H$\alpha$ filaments and their successive eruptions. In
the event, the interaction was initiated mainly by an active filament of them. They considered that the
second filament eruption may be the result of a loss of stability owing to the sudden mass injection
from the first filament eruption. \citet{liuy10}
presented another interesting case: two filaments erupted simultaneously and there was no transfer of
material between them during the initial stage; the two filaments merged together along the ejection path,
indicating the coalescence between the two interacting flux ropes. Moreover, \citet{bone09}
reported observations of the interaction and merging of two filaments, one active and one quiescent,
and their subsequent eruption. Even so, observations of interacting and erupting filaments are still rare at present.
With the multi-wavelength observations from Hinode, the Solar and Heliospheric Observatory (SOHO),
and the Solar Terrestrial Relations Observatory (STEREO), we investigate the interaction and eruption
of two filaments. We present observations of the filament interaction and mass ejection, and also measure
the bulk-flow velocities of the eruption using spectroscopic data. It is found that the two filaments
rise to approach each other, interact, and finally are ejected along two different paths. We describe the
observing instruments and data in Section \ref{obs}. The evolution of the filament interaction and
eruption is presented in Section \ref{picture}. Section \ref{erup_EIS} gives the bulk-flow velocities
of the eruption. Finally, a discussion is given in Section \ref{discussion}.
\section{Overview of the Observations and Instruments}
\label{obs}
The event of filament eruption was located in the solar active region NOAA 11045. The active region
produced several C-class and an M-class flares within a few hours on 2010 February 8. Among them,
a GOES C1.8 flare, which started at 11:00 UT and peaked at 11:14 UT, was in close association with
the filament eruption. Figure~\ref{goes_flux} shows the GOES 1--8 \AA~soft X-ray light curve around the flare time. In addition,
a white-light CME was first detected with SOHO/LASCO at 06:30 UT, having a linear speed of 153 km s$^{-1}$ and lasting about 12 hours.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\textwidth]{fig1.eps}
\end{center}
\caption{GOES 1--8 \AA~soft X-ray light curve for a period covering the filament interaction and eruption.
The vertical solid lines show the time range of the EIS scan over the active region,
while the dotted lines show the time period of filament eruption.}
\label{goes_flux}
\end{figure}
\subsection{Optical Images from Hinode/SOT}
The filament eruption was observed in H$\alpha$ images taken with Hinode/SOT (Solar Optical Telescope;
\citealt{tsun08}) with a spatial resolution of 0.16$''$ per pixel. SOT observed the event in
H$\alpha$ every half a minute before 10:22 UT and after 12:03 UT, but only obtained one image in between
(at 11:41 UT). During this period, SOT was observing the same active region using the Ca II H line with
a cadence of ten minutes; thus we can study the evolution of the two flare ribbons. SOT filtergraph
magnetograms are not available for this study. Instead, we use the data from SOHO/MDI (Michelson Doppler
Imager; \citealt{sche95}) to study the magnetic field in this active region.
\subsection{EUV Images from SOHO/EIT and STEREO/SECCHI/EUVI}
The interaction and eruption of the filaments were observed in EUV by SOHO/EIT (Extreme-ultraviolet
Imaging Telescope; \citealt{dela95}) and by the SECCHI/EUVI (Sun Earth Connection Coronal and Heliospheric
Investigation/Extreme UltraViolet Imager; \citealt{wuls04}) instrument onboard the STEREO spacecraft
(A and B). The EIT images were obtained in 195 \AA~($\sim$1.5$\times$10$^6$~K) with a cadence of 12 minutes,
while the SECCHI/EUVI images were observed in 304 \AA~($\sim$6.0$\times$10$^4$~K) and 195 \AA~with
cadences of 10 and 5 minutes, respectively. The active region was located near the center of the solar
disk in the EIT field of view (FOV), while it appeared on the east and west limbs in the STEREO-A and B FOVs,
respectively. On that day, these two spacecraft were separated by an angle of $\sim$136$^{\circ}$. We
use the EUV data between 10:40 UT and 11:30 UT. This period is from about 20 minutes prior to the flare
onset to about 30 minutes after it, thus covering the entire eruption process.
\subsection{Spectroscopic Data from Hinode/EIS}
We measure the Doppler velocities of the mass ejection from spectroscopic data taken with Hinode/EIS
(Extreme-ultraviolet Imaging Spectrometer; \citealt{culh07}). EIS covers two wavelength bands, 170--211
\AA~and 246--292 \AA, referred to as the short and long wavelength bands, respectively. These bands
include some strong transition region and coronal lines over a wide temperature range. We mainly analyze
the Fe {\sc xii} 195.12 \AA~line here. In observations, the 2$''$ slit of EIS was used to scan
the active region of 324$''\times$376$''$ with an exposure time of 45 s. The scanning
lasted for 2 hours and 6 minutes, as indicated in Figure \ref{goes_flux}.
Our study focuses on the region hosting the eruption, which has a size of about
30$''\times$90$''$. EIS slit crossed this region from 11:11 UT to 11:22 UT
(between the two vertical dotted lines in Figure \ref{goes_flux}). Note that, for all the images from EIS,
we use the coordinates in arcseconds relative to a reference point inside the EIS raster FOV.
We reduce the EIS data using the standard processing package. This corrects the detector bias and dark
current, as well as hot pixels and cosmic ray hits, resulting in absolute intensities in
erg~cm$^{-2}$~s$^{-1}$~sr$^{-1}$~\AA$^{-1}$. We also make a correction for a slight tilt of the slit
on the CCDs. In addition, we correct for the variation in spectral line positions over the Hinode
orbit caused by temperature variations in the spectrometer. Such an orbital variation is obtained by
averaging the centroid positions over the length of the slit for a quiet region. When measuring the
relative Doppler velocity, we use the average line center over a quiet region as the reference Fe {\sc xii}
wavelength.
To interpret the multi-wavelength observations, it is important to co-align the different images.
There is an instrumental offset between the images taken in the two EIS CCDs \citep{youn07a}.
To correct this offset, we shift the long wavelength images by 2$''$ in the solar X-direction
and 17$''$ in the solar Y-direction. To co-align the images from SOT and MDI, we use the
sunspots observed by both instruments. The co-alignment between the EIT and EIS data
is made using the 195 \AA~images obtained by both instruments. We estimate the uncertainty
of the co-alignment to be within 5$''$--10$''$.
\section{Filament Interaction and Eruption}
\label{picture}
\subsection{The Filaments and Associated Flare}
\label{ejec_SOT}
Figure \ref{Ha_MDI} shows the SOT H$\alpha$ images of the flaring region overlaid with the contours
of the MDI magnetogram. The two H$\alpha$ images were observed $\sim$40 minutes prior to and after
the flare onset at 11:00 UT, respectively. The superimposed MDI magnetogram was taken at 11:15 UT,
with contour levels of 200, 500, 900, and 1500 G. There are some strong positive (red contours)
and negative (blue contours) magnetic patches. The white arrows (Figure \ref{Ha_MDI}a) mark the
two filaments prior to the flare onset.
Considering the filament morphology and the magnetic configuration, it is reasonable to speculate
that the dark segments indicated by the two left arrows are magnetically connected with each other, or
that they are the two segments of one large filament. In this context, we only mention one filament,
which includes the two segments.
After the flare (Figure \ref{Ha_MDI}b), the two filaments disappeared, and some post-flare loops
(shown by the yellow arrow) appeared instead. From the two H$\alpha$ images, we find that the filaments
erupted between 10:22 UT and 11:41 UT.
The interaction between the left and right filaments and their eruption
are more clearly presented in the EUV images described in Section \ref{erup_EUV}.
\begin{figure}
\centerline{\includegraphics[width=1.0\textwidth]{fig2.eps} }
\caption{Hinode/SOT~H$\alpha$ images before and after the flare onset at 11:00 UT,
superimposed with the MDI magnetogram taken at 11:15 UT on February 8. Red and blue
contours refer to the positive and negative polarities, respectively, with levels of
200, 500, 900, and 1500 G. The white arrows in panel (a) point to the two close filaments
existing prior to the flare onset. The yellow arrow in panel (b) marks the post-flare loops.}
\label{Ha_MDI}
\end{figure}
Figure \ref{CaIIH} shows the evolution of the two flare ribbons. The SOT Ca II H images have the
same FOV as the H$\alpha$ images in Figure \ref{Ha_MDI}. The left filament was located
between the two flare ribbons, and the right one was on the right side of the ribbons.
The two flare ribbons began to brighten
at $\sim$11:00 UT (Figure \ref{CaIIH}c). They became the brightest near the peak time of the flare soft X-ray emission
(Figure \ref{CaIIH}d). Tens of minutes later, the ribbons gradually disappeared (Figures \ref{CaIIH}e
and \ref{CaIIH}f).
\begin{figure}
\centerline{\includegraphics[width=1.0\textwidth]{fig3.eps} }
\caption{SOT~Ca II H images during the flare. The image in panel (a) is overlaid with
the MDI magnetogram taken at 11:15 UT. The MDI contours and the FOV are the same as
that in Figure \ref{Ha_MDI}.}
\label{CaIIH}
\end{figure}
\subsection{The Filament Interaction and Eruption}
\label{erup_EUV}
The filament interaction and eruption are most evident in EUV images from EIT
and SECCHI/EUVI. In the FOVs of these instruments, the eruptions appeared near the center
of the solar disk and on the east and west limbs, respectively, as mentioned above. Therefore,
we can view this event in three different perspectives. Figures \ref{EUVI304} and \ref{195all}
give the pictures in the filters of 304 \AA~and 195 \AA. Prior to the onset of the flare,
we can see the loop expansion due to the filament rising. There appeared some bright loop
structures (marked by the black arrows in Figures \ref{EUVI304}a, \ref{EUVI304}b, \ref{195all}a,
\ref{195all}c, and \ref{195all}e) at $\sim$10:56 UT below the core region as compared with
the structure at an earlier time (Figure \ref{195all}b). In particular, there were some loops
being contacted (Figure \ref{195all}e). This may lead to or accelerate the filament interaction
and eruption.
The two EUV filament channels (EFCs) can be seen in Figures \ref{EUVI304}a, \ref{EUVI304}b,
\ref{195all}a, \ref{195all}c, and \ref{195all}d, as marked by the white arrows.
At an early time ($\sim$10:56 UT), both of the filaments were relatively stable.
They were rising higher and about to erupt at $\sim$11:06 UT;
meanwhile, the two flare ribbons were brightened (see Figure \ref{CaIIH}).
A few minutes later, the two filaments collided and began to erupt,
as shown in Figure \ref{195all}i. Due to the low temporal and spatial resolutions, we
cannot distinguish which segment (or both segments) of the left filament met with
the right filament. However, we can judge that the right EFC made a
collision with the left EFC. This was also confirmed by the observation
that the ejected material from the filament suddenly changed
the direction of ejection: the cool material was expelled to
the left side as shown in Figure \ref{195all}h when the two EFCs came into contact
with each other (shown more clearly in the video 1, available in the electronic edition
of the paper). It seems that the two filaments erupted nearly simultaneously
after the peak of the flare soft X-ray emission at 11:14 UT (Figures \ref{EUVI304}e--h, \ref{195all}g, and
\ref{195all}l). Moreover, a careful inspection reveals two ejection
paths, marked by the two arrows in Figures \ref{EUVI304}e, \ref{EUVI304}h, \ref{195all}g,
and \ref{195all}l. Because of the projection effect, the two ejection paths are not clearly visible
in EIT images (Figure \ref{195all}k). The eruptions along different directions
can be seen more clearly in the movies (videos 2 and 3). At the later stage of the eruption, some of
the filament material cooled down (Figure \ref{195all}j) and fell back to the solar surface,
while the remainder was ejected into the space, as judged from the ejection velocities measured
in Section \ref{erup_EIS}.
\begin{figure}
\centerline{\includegraphics[width=0.7\textwidth]{fig4.eps} }
\caption{SECCHI/EUVI 304~\AA~images of the filament eruption region (evolving from top to bottom).
``SA'' refers to STEREO-A and ``SB'' to STEREO-B. The black arrows in (a) and (b) mark the bright
expanding loops. The white arrows point to the filaments or EFCs.}
\label{EUVI304}
\end{figure}
\begin{figure}
\centerline{\includegraphics[width=1.0\textwidth]{fig5.eps} }
\caption{EIT 195 \AA~images and EUVI 195 \AA~images of the filament eruption region (evolving from top to bottom).
``SA'' and ``SB'' refer to STEREO-A and STEREO-B, respectively. In (b), the EIT image
is overlaid with the MDI magnetogram taken at 11:15 UT. The black arrows in (a), (c),
and (e) mark the bright expanding loops. The white arrows point to the filaments or EFCs.
In (k), the white box shows the filament eruption region that we focus on in the study.}
\label{195all}
\end{figure}
From different view angles (Figure \ref{195all}), we see the different manifestations of
filament eruption. Near the solar disk center, the filament eruption appears rather diffuse; while
at the limbs the observed ejection is rather collimated. Furthermore, we observe the different
shapes of the erupting filaments from different passbands. In the 304 \AA~image, the filament
ejection is more extended; while in the 195 \AA~image, the ejected material is more confined.
The ejection being visible in emission in a few passbands suggests that the filament material
is composed of plasmas with a wide temperature range from 6.0$\times$10$^4$~K to 1.5$\times$10$^6$~K
(see also Figure \ref{EISinten}). It is possible that part of the filament material is heated to the
coronal temperature during the interaction and eruption. An alternative view is that the ejection
contains hot plasmas from the reconnection outflow. In a word, cool material ($\sim$10$^{4}$~K) and
hot material ($\sim$10$^{6}$~K) co-exist in the eruption.
\section{Bulk-flow Velocities of the Eruptions}
\label{erup_EIS}
The EIS spectroscopic data were obtained by scanning the active region from solar west to east
between 10:10 UT and 12:16 UT.
Figure \ref{EISinten} shows the intensity maps in EIS spectral lines. The white boxes mark the
main eruption region, which was scanned from 11:11 UT to 11:22 UT. During that time, the filaments
were erupting. We can see the bright emission from the flare kernel and from the
filament eruption in most spectral lines. The filament material is visible in the low temperature line
of Mg {\sc v}~($\sim$3.0$\times$10$^5$~K), as well as in the high temperature line of
Fe {\sc xv}~($\sim$2.0$\times$10$^6$~K).
\begin{figure}
\centerline{\includegraphics[width=1.0\textwidth]{fig6.eps} }
\caption{Intensity maps of the EIS spectral lines. The white boxes mark the main eruption
region with the same FOV as the white box in Figure \ref{195all}. The filament material is visible
in most of these images. Note that the coordinates are relative to a reference point in
arcseconds inside the EIS raster FOV (the same for Figures \ref{EIS195} and \ref{EIS195part}),
and that there is an instrumental offset between the two EIS CCDs recording the short and
long wavelength images, respectively. The dark/bright vertical lines refer to missing/abnormal
data points.}
\label{EISinten}
\end{figure}
Figure \ref{EIS195} shows the Fe {\sc xii} 195.12 \AA~intensity and Doppler velocity maps.
In this figure, the Doppler velocity is derived from a single Gaussian fitting to the observed
spectral line. We detect significant outflows (negative velocities) with a speed of hundreds
of km s$^{-1}$ in the eruption region. In addition, the analysis also reveals some weak redshift
(positive velocities) of about tens of km s$^{-1}$ near the edge of the region (indicated by the arrows).
The redshifts may originate from the cooling plasmas, which were falling back to the solar surface
at the late stage of the eruption when the EIS slit passed over the edge of the
region. When measuring the Doppler velocity, we notice that, in the quiet and redshifted regions,
the line profiles can be well fitted by a single Gaussian function; while in the region with
significant outflows, the line profiles show obvious double Gaussian components, with a blueshifted
component superimposed on a relatively static one. Therefore, we consider to use the double Gaussian
function to fit the line profiles in the main eruption region. Figure \ref{EIS195part} shows
the results of the Fe {\sc xii} line. In the velocity panel, most of the numbers are
derived from a double Gaussian fitting. We also calculate the ratio \textit{r},
of the integrated intensity of the blueshifted component to that of the entire profile.
It is plotted in the ratio panel (\textit{r} becomes zero when the line profile is well
fitted by a single Gaussian function). We find that at many locations the blueshifted
component is stronger than the static component, i.e., the intensity ratio as defined above
is over 0.5. Figure \ref{profile} shows an example of double Gaussian fitting for the pixel
marked by the plus sign in Figure \ref{EIS195part}. The intensity of the blueshifted component
reaches even 67\% of the total intensity, and the blueshifted velocity is 195 km s$^{-1}$.
Double Gaussian shapes are also present in other line profiles (such as Fe {\sc xiii} and Fe {\sc xv}).
In the region marked by the arrow in the velocity panel of Figure \ref{EIS195part}, the velocity
of the blueshifted component can be as high as 300 km s$^{-1}$, though its intensity is less
than half of the total. If corrected for the projection effect, the true velocity may be
greater than the escape velocity on solar surface ($\sim$618 km s$^{-1}$).
It means that there is some material
going into the space as mentioned in Section \ref{erup_EUV}. In fact, the time when the EIS
slit scanned over this region, $\sim$11:11 UT, corresponds to the onset of the filament eruption.
\begin{figure}
\centerline{\includegraphics[width=1.0\textwidth]{fig7.eps} }
\caption{EIS 195~\AA~intensity and Doppler velocity maps. The white solid box marks the
filament eruption region. The white dashed box represents the quiet region that we use to
calculate the reference wavelength. All the values in the velocity panel are obtained
from the single Gaussian fitting. The arrows point to the downflow areas at the edge
of the main eruption region.}
\label{EIS195}
\end{figure}
\begin{figure}
\centerline{\includegraphics[width=1.0\textwidth]{fig8.eps} }
\caption{EIS 195~\AA~intensity and Doppler velocity maps of the filament eruption region
marked by the white solid box in Figure \ref{EIS195}. The velocity field is obtained by
combining the results from either the double Gaussian fitting or the single Gaussian fitting.
When the double Gaussian fitting is applied, the velocities refer to those of the blueshifted
components. The right panel shows the intensity ratio between the blueshifted component and
the whole profile. The ratio equals zero when the profile can be well fitted by a single
Gaussian function. The white arrow points to the region showing large Doppler velocities over
250 km s$^{-1}$ at the beginning of the eruption. The line profile and its fitting result
at the plus sign is shown in Figure \ref{profile}.}
\label{EIS195part}
\end{figure}
\begin{figure}
\centerline{\includegraphics[width=0.8\textwidth]{fig9.eps} }
\caption{A typical line profile and its double Gaussian fitting at the plus sign in Figure
\ref{EIS195part}. The histogram is the observed profile and the solid line is the fitting
result. The vertical dotted line refers to the reference wavelength. This line is well
fitted by double Gaussian components that are plotted with dashed curves. The velocity
for the blueshifted component is 195 km s$^{-1}$, and its intensity ratio to the whole
profile is 0.67.}
\label{profile}
\end{figure}
The spectral line profile in the eruption region cannot be fitted by a single Gaussian function.
Although the low spatial resolution of the scanning observation may cause such a deviation from Gaussian,
we think it more likely that there are multi-velocity components along the line of sight.
This is supported by the observation that the two filaments are ejected along different trajectories.
\section{Discussion}
\label{discussion}
We study the interaction and eruption of two filaments observed from three different view angles.
The event was accompanied by a C1.8 two-ribbon flare. At first, we observe the expansion of coronal loops.
Then, the two filaments rose higher, interacted with each other, and were finally ejected
along two different paths. During the interaction and eruption, the flare ribbons were brightened
and then gradually decay. The filament eruption was observed in emission in both the 304 \AA~image and the 195 \AA~band,
suggesting that the erupted filament material contained plasmas with a wide temperature range
of 10$^{4}$--10$^{6}$ K. We measure the bulk-flow velocities using the EIS spectroscopic data. Significant
outflows were detected with a speed of several hundred km s$^{-1}$ during the eruption. Downflows of tens
of km s$^{-1}$ were also observed at the edge of the eruption region during the late stage of the eruption.
Most of the blueshifted line profiles are double-peaked and can be fitted to two gaussians. After correcting
for the projection effect, we find that the upward velocity may exceed the solar
escape velocity, suggesting that some filament material may be ejected into the space.
This event occurred in an active region that produced a number of large flares. From
the movie of the MDI 96-min magnetogram (video 4, available in the electronic edition), we can
see that the strong positive and negative magnetic patches were approaching, disintegrating,
canceling, and disappearing, especially in the flaring region at $\sim$11:00 UT.
From the available data we find that there is a close relationship between the flare and the
filament eruption. Before the flare started, the coronal loops appeared to expand. Then the filaments
rose higher. Several minutes later, the flare ribbons were brightened.
It is therefore likely that magnetic reconnection occurred between the two filaments and triggered the filament eruption.
As the filaments rose, they began to be linked together, interact with each other, and were finally erupted. After the
peak of the flare, the filaments were ejected along two different paths. Some of the material escaped
from the Sun, and the remainder cooled down and fell back to the solar surface. The interaction of the
two filaments may lead to or accelerate the flare energy release.
It has been discussed that pairs of filaments (or filament segments) can merge and interact with each other
\citep{daza48, mart98, martens01, schm04, devo05, aula06}. However, these studies are mainly focused on
the filament formation. Observations of the evolution and eruption of filament pairs are
valuable and can provide a better understanding of the filament nature and coronal dynamics.
To our knowledge, \citet{sujt07} reported the first observational study of the interaction
between two distinct H$\alpha$ filaments and the successive eruptions. In their event, the two
filaments erupted with an interval of 40 minutes. It is observed that some material was transferred
from the filament that erupted first to the other filament, and triggered its eruption.
This event was associated with a CME. Quite recently, \citet{liuy10} presented another
interesting event: two filaments erupted simultaneously and there was no mass transfer
between them during the initial stage. The two filaments merged together along the ejection
path and no CME was observed. Furthermore, \citet{bone09} reported observations of
the interaction and full merging of two filaments, one active and the other quiescent, and their
subsequent eruption. In these observational studies, the authors mainly used
the ground-based H$\alpha$ data with high cadence that observes filament plasmas at a low temperature.
In our study, however, we use multi-wavelength images in both EUV and H$\alpha$ to show
the interaction and eruption of the two filaments from three different view angles. We can
compare this event with the previously reported events. The two filaments in this event seem to erupt
simultaneously without evident mass transfer between them. This is similar to the case of
\citet{liuy10}, but different from the one studied by \citet{sujt07}.
Moreover, our event was associated with two active filaments, while the event presented by
\citet{bone09} involved one active filament and one quiescent filament. In particular, we find
that the two filaments were ejected along two different paths. These different paths of ejections
can be explained by the 3D magnetic reconnection as shown in numerical simulations of
\citet{jiang11a, jiang11b}. The magnetic reconnection occurred when the two filaments were rising and interacting with each other.
Moreover, in observations, the EFCs did not seem to merge; therefore,
they were linked only partially before the filament material was ejected. This may also have some relation
to the different ejection paths. The details need to be verified by numerical simulations.
The chirality (handedness) of filaments plays an important role in the
merging and linkage of filament pairs. Our results support the scenario that
only filaments of the same chirality can link up \citep{martens01, schm04, devo05, aula06,sujt07}.
In our event, the two filaments are both dextral, and they are linked up and interact with each other. This
is consistent with the previous research.
Spectroscopic data are important for the study of filament heating and coronal dynamics \citep{ying09}. In this event,
we only analyze the EIS Fe {\sc xii} 195.12 \AA~line to obtain the bulk-flow velocity of the filament eruption.
In fact, the EIS line profiles are complicated in most regions, especially in eruption regions.
They are often double-peaked, consisting of static and shifted components \citep{ying11}, which
can be well fitted by double Gaussians. The double-peaked profile may indicate presence of multiple line-of-sight
velocities or may result from the low spatial resolution of the observation.
In either case, the dynamics of the filaments and corona are very complicated during the eruption.
In addition, we can use the spectroscopic data to measure the temperature and density of filaments and
to analyze the plasma heating and cooling during the eruption. To this end, we need spectroscopic
data and multi-wavelength images with higher temporal and spatial resolutions.
\begin{acknowledgements}
The authors would like to thank P.~F. Chen, J. Qiu, and Y. Guo for their helpful discussions
and the referee for valuable comments on the paper. This work
was supported by NSFC under grants 10828306 and 10933003 and by NKBRSF under grant 2011CB811402.
Hinode is a Japanese mission developed and launched by ISAS/JAXA, collaborating with NAOJ as a
domestic partner, and NASA (USA) and STFC (UK) as international partners. Scientific operation
of the Hinode mission is conducted by the Hinode science team organized at ISAS/JAXA. Support for
the post-launch operation is provided by JAXA and NAOJ (Japan), STFC (U.K.), NASA, ESA, and NSC
(Norway).
\end{acknowledgements}
|
\section{Introduction}
The lightbulb process introduced by \cite{rrz} was motivated by a pharmaceutical study of the effect
of dermal patches designed to activate targeted receptors.
An active receptor will become inactive, and an inactive one active,
if it receives a dose of medicine released from the dermal patch. On each of $n$ successive
days $r=1,\ldots,n$ of the study, exactly $r$ randomly selected receptors will each receive one dose
of medicine from the patch, thus changing, or toggling, their status between the active and inactive states.
We adopt the more colorful language of ~\cite{rrz}, where receptors are represented by lightbulbs that
are being toggled between their on and off states.
Some fundamental properties of $W_n$, the number of light bulbs on at the end of day $n$, were derived in \cite{rrz}. For instance, Proposition~2 of \cite{rrz} shows that when $n(n+1)/2 = 0 \,\mbox{mod}\,2$, or, equivalently, when $n\,\mbox{mod}\,4 \in \{0,3\}$, the support of $W_n$ is a subset of even integers up to $n$, and that otherwise the support of $W_n$ is a set of odd integers up to $n$.
Further, in \cite{rrz}, the mean and variance of $W_n$ were computed, and based on numerical computations, an approximation of the distribution of $W_n$ by the `clubbed' binomial distribution was suggested.
To describe the clubbed binomial, let $Z_n$ be a binomial $\mbox{Bin}(n-1,1/2)$ random variable, and for $i \in \mathbb{Z}$ let $\pi_i^*=P(Z_n=i)$, that is
\begin{eqnarray*}
\pi_i^* = \left\{
\begin{array}{cc}
{n-1 \choose i} \left(\frac{1}{2}\right)^{n-1} & \mbox{for $i=0,1,\ldots,n-1$,}\\
0 & \mbox{otherwise.}
\end{array}
\right.
\end{eqnarray*}
Let $L_{1,n}$ and $L_{0,n}$ denote the set of all odd and even numbers in $\{0,1,\ldots,n\}$, respectively.
Define, for $m=0,1$,
$$\pi^m_i=\left\{\begin{array}{ll}
\pi_{i-1}^*+\pi_i^*,&\ i\in L_{m,n},\\
0,&\ i\not\in L_{m,n}.
\end{array}\right.
$$
Summing binomial coefficients using `Pascal's triangle' yields
\begin{eqnarray} \label{pascal}
\pi^m_i=\left\{\begin{array}{ll}
{n\choose i}\left(\frac{1}{2}\right)^{n-1},&\ i\in L_{m,n},\\
0,&\ i\not\in L_{m,n}.
\end{array}\right.
\end{eqnarray}
We say that the random variable $C_{m,n}$ has the clubbed binomial distribution if $P(C_{m,n}=i)=\pi^m_i$ for $i \in L_{m,n}$. In words, the clubbed binomial distribution is formed by combining two adjacent cells of
the binomial.
It was observed in ~\cite{rrz} that the clubbed binomial distribution appeared to approximate the
lightbulb distribution $W_n$ exponentially well. Here we make that observation rigorous by supplying an exponentially decaying bound in total variation. First, recall that if $X$ and $Y$ are two random variables with distributions supported on $\mathbb{Z}$, then the total variation distance between the (laws of) $X$ and $Y$, denoted $d_{{\mbox{\scriptsize \rm TV}}}(X,Y)$, is given by
\begin{eqnarray} \label{def:dtv}
d_{{\mbox{\scriptsize \rm TV}}}(X,Y)=\sup_{A \subset \mathbb{Z}}|P(X \in A)-P( Y \in A)|.
\end{eqnarray}
\begin{theorem}
\label{thm:illuminate}
Let $W_n$ be the total number of bulbs on at the terminal time in the lightbulb process of size $n$ and let $C_n=C_{m,n}$
where $m=0$ for $n\,{\rm mod}\,4\in\{0,3\}$ and $m=1$ for $n\,{\rm mod}\,4\in\{1,2\}$. Then
\begin{eqnarray*}
d_{{\mbox{\scriptsize \rm TV}}}(W_n,C_n) \le 2.7314 \sqrt{n} e^{-(n+1)/3}.
\end{eqnarray*}
\end{theorem}
In particular, the approximation error is less than 1\% for $n\ge 21$ and less than 0.1\% for $n\ge 28$.
A Berry-Esseen bound in the Kolmogorov metric of order $1/\sqrt{n}$ for the distance between the standardized value of $W_n$ and the unit normal was derived in ~\cite{gz}. The lighbulb chain was also studied in ~\cite{zl}, and served there as a basis for the exploration of the more general class of Markov chains of multinomial type. One feature of such chains is their easily obtainable spectral decomposition, which informed the analysis in \cite{gz}. In contrast, here we demonstrate the exponential bound in total variation using only simple properties of the lightbulb process.
After formalizing the framework for the lightbulb process in the next section, we prove Theorem \ref{thm:illuminate} by Stein's method. In particular, we develop a Stein operator ${\cal A}$ for the clubbed binomial distribution and obtain bounds on the solution $f$ of the associated Stein equation. The exponentially small distance between $W_n$ and the clubbed binomial $C_n$ can then be seen to be a consequence of the vanishing of the expectation of ${\cal A}f$ except on a set of exponentially small probability.
\section{The lightbulb process}
We now more formally describe the lightbulb process. With $n \in
\mathbb{N}$ fixed we will let ${\bf X}=\{X_{rk}: r=0,1,\ldots,n ,k=1,\ldots,n\}$ denote a collection of
Bernoulli variables. For $r \ge 1$ these `switch' or `toggle' variables
have the interpretation that
\begin{eqnarray*} X_{rk} &=& \left\{
\begin{array}{cc}
1 & \mbox{ if the status of bulb $k$ is changed at stage $r$,} \\
0 &\mbox{ otherwise.}
\end{array}
\right.
\end{eqnarray*}
We take the initial state of the bulbs to be given deterministically by setting the switch variables $\{X_{0k},k=1,\ldots,n\}$ equal to zero, that is, all bulbs begin in the off position.
At stage $r$ for $r=1,\ldots,n$, $r$ of the $n$ bulbs are chosen uniformly to have their status changed, with
different stages mutually independent. Hence, with $e_1,\ldots,e_n \in \{0,1\}$,
the joint distribution of $X_{r1},\ldots,
X_{rn}$ is given by
\begin{eqnarray*}
\lefteqn{P(X_{r1} = e_1, \cdots, X_{rn}=e_n)
=\left\{
\begin{array}{cc}
{n \choose r}^{-1} & \mbox{if $e_1+\cdots+e_n = r,$} \\
0 & \mbox{ otherwise,}
\end{array}
\right.}\\
&& \nonumber \mbox{with the collections $\{X_{r1},\ldots,X_{rn}\}$
independent for $r=1,\ldots,n$.}
\end{eqnarray*}
Clearly, at each stage $r$
the variables $(X_{r1},\cdots, X_{rn})$ are exchangeable.
For $r,i = 1, \ldots, n$, the
quantity $\left(\sum_{s=1}^r X_{si}\right) \mbox{ mod }2$ is the
indicator that bulb $i$ is on at time $r$ of the lightbulb process, so letting
\begin{eqnarray*}
I_i = \left(\sum_{r=0}^n X_{ri}\right) \mbox{ mod }2
\qmq{and} W_n=\sum_{i=1}^n I_i,
\end{eqnarray*}
the variable $I_i$ is the
indicator that bulb $i$ is on at the terminal time, and $W_n$ is the number of bulbs on at the terminal time.
The lightbulb process is a special case of a class of multivariate chains
studied in \cite{zl}, where randomly chosen subsets of $n$ individual particles evolve according
to the same marginal Markov chain. As shown in ~\cite{zl},
such chains admit explicit full spectral
decompositions, and in particular, the transition matrices for each stage of the lightbulb
process can be simultaneously diagonalized by a Hadamard matrix. These properties
were applied in \cite{rrz} for the calculation of the
moments needed to compute the mean and variance of $W_n$ and to develop
recursions for the exact distribution, and in \cite{gz} for a Berry-Esseen bound of
the standardized $W_n$ to the normal.
\section{Stein Operator}
In order to apply Stein's method, we first develop a Stein equation for the clubbed binomial distribution $C_{m,n}$ and then present bounds on its solution. With $\pi^m_x$ given by (\ref{pascal}), let
$\pi^m(A)=\sum_{x \in A} \pi^m_x$. Set $\alpha_x=(n-x)(n-1-x)$ and $\beta_x=x(x-1)$ for $x\in\{0,\dots,n\}$. One may easily directly verify the balance equation
\begin{eqnarray} \label{eqn:balance}
\alpha_{x-2}\pi^m_{x-2} = \beta_x \pi^m_x \quad \mbox{for $x \in L_{m,n}$},
\end{eqnarray}
which gives the generator of the distribution
of $C_{m,n}$ as
\begin{eqnarray} \label{generator}
{\cal A}f(x)=\alpha_x f(x+2)-\beta_x f(x),\mbox{ for } x\in L_{m,n}.
\end{eqnarray}
For $A \subset L_{m,n}$, we consider the Stein equation
\begin{eqnarray} \label{Afr}
{\cal A}f_A(x)=1_{A}(x)-\pi^m(A),\ x\in L_{m,n}.
\end{eqnarray}
For a function $g$ with domain $A$ let $\|g\|$ denote $\sup_{x \in A} |g(x)|$.
\begin{lemma}
\label{lem:bound}
For $m\in\{0,1\}$ and $A=\{r\}$ with $r \in L_{m,n}$, the unique solution $f^m_r(x)$ of (\ref{Afr}) on $L_{m,n}$ satisfying the boundary condition $f^m_r(m)=0$ is given, for $m<x\le n, x \in L_{m,n}$, by
\begin{eqnarray} \label{frsol}
f^m_r(x)= \left\{
\begin{array}{cl}
-\frac{\pi^m([0,x-2]\cap L_{m,n})\pi^m_r}{\beta_x \pi^m_x} & \mbox{for $m < x<r+2$} \\
\frac{\pi^m([x,n]\cap L_{m,n})\pi^m_r}{\beta_x \pi^m_x} & \mbox{for $r+2 \le x \le n$.}
\end{array}
\right.
\end{eqnarray}
Furthermore, for all $A \subset L_{m,n}$, $f^m_A(x)=\sum_{r\in A}f^m_r(x)$ is a solution of (\ref{Afr}) and satisfies
\begin{eqnarray*}
\|f^m_A\| \le \frac{2.7314}{\sqrt{n}(n-1)} \quad \mbox{for $n \ge 1$.}
\end{eqnarray*}
\end{lemma}
Lemma \ref{lem:bound} is proved in Section \ref{sec:bound}.
Applying Lemma \ref{lem:bound}, we now prove our main result.
{\em Proof of Theorem \ref{thm:illuminate}:} Fix $m\in\{0,1\}$ and $A \subset L_{m,n}$, and let $f:=f^m_A$ be the solution to (\ref{Afr}). Dropping subscripts,
let $W=\sum_{i=1}^n I_i$, where $I_i$ is the indicator that bulb $i$ is on at the terminal time. For $i,j \in \{1,\ldots,n\}$, now with slight abuse of notation, let $W_i=W-I_i$, and for $i \not = j$ set $W_{ij}=W-I_i-I_j$.
Then
\begin{eqnarray*}
\lefteqn{E(n-W)(n-1-W)f(W+2)}\\
&=&E\sum_{i=1}^n (1-I_i)(n-1-W)f(W_i+2)\\
&=&E\sum_{i \not =j} (1-I_i)(1-I_j)f(W_{ij}+2),
\end{eqnarray*}
and similarly,
\begin{eqnarray*}
E W(W-1)f(W) = E\sum_{i=1}^n I_i W_i f(W_i+1) = E\sum_{i \not = j} I_i I_j f(W_i+1)= E\sum_{i \not = j} I_i I_j f(W_{ij}+2).
\end{eqnarray*}
By Proposition~2 of \cite{rrz}, $P(W \in L_{m,n})=1$,
and hence (\ref{Afr}) holds upon replacing $x$ by $W$. Taking expectation and using the expression for the generator in (\ref{generator}), we obtain
\begin{eqnarray} \label{gen-indicators}
P(W \in A) - \pi^m(A)
= E{\cal A}f(W)= E \sum_{i \not = j} \left( (1-I_i)(1-I_j)-I_iI_j \right) f(W_{ij}+2).
\end{eqnarray}
Recalling that $X_{rk}$ is the value of the switch variable at time $r$ for bulb $k$, let $A_{ij}$ be the event that the switch variables of the distinct bulbs $i$ and $j$ differ in at least one stage, that is, let
\begin{eqnarray} \label{def:Aij}
A_{ij}= \bigcup_{r=1}^n \{X_{ri} \not = X_{rj}\}.
\end{eqnarray}
Now using (\ref{gen-indicators}) we obtain
\begin{eqnarray}
&& \left| P(W \in A) - \pi^m(A) \right| = \left| E \sum_{i \not = j} \left( (1-I_i)(1-I_j)-I_iI_j \right) f(W_{ij}+2) \right| \nonumber \\
&\le& \left| \sum_{i \not = j} E \left( (1-I_i)(1-I_j)-I_iI_j \right) f(W_{ij}+2){\bf 1}_{A_{ij}} \right| \nonumber \\
&& + \left| \sum_{i \not = j} E \left( (1-I_i)(1-I_j)-I_iI_j \right)f(W_{ij}+2){\bf 1}_{A_{ij}^c} \right|. \label{twoterms}
\end{eqnarray}
Note that $I_i, I_j \in \{0,1\}$ implies
$$
(1-I_i)(1-I_j){\bf 1}_{I_i \not = I_j}=0=I_iI_j {\bf 1}_{I_i \not = I_j},
$$
and hence for the first term in (\ref{twoterms}) we obtain
\begin{eqnarray}
\lefteqn{\sum_{i \not = j} \left( (1-I_i)(1-I_j)-I_iI_j \right) f(W_{ij}+2) {\bf 1}_{A_{ij}}}\nonumber \\
&=& \sum_{i \not = j} \left( (1-I_i)(1-I_j)-I_iI_j \right) f(W_{ij}+2){\bf 1}_{A_{ij},I_i=I_j}. \label{eq:fterm}
\end{eqnarray}
For a given pair $i,j$, on the event $A_{ij}$ let $t$ be any index for which $X_{ti} \not = X_{tj}$, and let ${\bf X}^{ij}$ be the collection of switch variables given by
\begin{eqnarray*}
X_{rk}^{ij}= \left\{
\begin{array}{cl}
X_{rk} & r \not = t,\\
X_{tk} & r=t, k \not \in \{i,j\},\\
X_{ti} & r=t, k=j,\\
X_{tj} & r=t, k=i.
\end{array}
\right.
\end{eqnarray*}
In other words, in stage $t$, the unequal switch variables $X_{ti}$ and $X_{tj}$ are interchanged, and all other
variables are left unchanged. Let $I_k^{ij}$ be the status of bulb $k$ at the terminal time when applying switch variables ${\bf X}^{ij}$, and similarly set
$W_{ij}^{ij}=\sum_{k \not \in \{i,j\}} I_k^{ij}$. Note that as the status of both bulbs $i$ and $j$ are toggled upon interchanging their stage $t$ switch variables, and all other variables are unaffected, we obtain
$$
I_i^{ij}=1-I_i, \quad I_j^{ij}=1-I_j \qmq{and} W_{ij}^{ij}=W_{ij}.
$$
In particular, $I_i=I_j$ if and only if $I_i^{ij}=I_j^{ij}$, and, with $A_{ij}^{ij}$ as in
(\ref{def:Aij}) with $X_{rk}^{ij}$ replacing $X_{rk}$, we have additionally that $A_{ij}^{ij}=A_{ij}$. Further, by exchangeability
we have ${\cal L}({\bf X})={\cal L}({\bf X}^{ij})$. Therefore,
\begin{eqnarray*}
\lefteqn{E (1-I_i)(1-I_j)f(W_{ij}+2){\bf 1}_{A_{ij},I_i=I_j}}\\
&=&E (1-I_i^{ij})(1-I_j^{ij})f(W_{ij}^{ij}+2){\bf 1}_{A_{ij}^{ij},I_i^{ij}=I_j^{ij}}\\
&=&E I_i I_j f(W_{ij}+2){\bf 1}_{A_{ij},I_i=I_j},
\end{eqnarray*}
showing, by (\ref{eq:fterm}), that the first term in (\ref{twoterms}) is zero. Therefore,
\begin{eqnarray*}
\lefteqn{|P(W \in A) - \pi^m(A) |}\\
&&\le \left| \sum_{i \not = j} E \left( (1-I_i)(1-I_j)-I_iI_j \right) f(W_{ij}+2){\bf 1}_{A_{ij}^c} \right|
\le \|f\| \sum_{i \not = j}P(A_{ij}^c).
\end{eqnarray*}
As $A_{ij}^c$ is the event that the switch variables of $i$ and $j$ are equal in every stage, recalling that
these variables are independent over stages we obtain
\begin{eqnarray*}
P(A_{ij}^c) &=& \prod_{r=1}^n \frac{r(r-1)+(n-r)(n-1-r)}{n(n-1)}\\
&=& \prod_{r=1}^n \left( 1 - \frac{2(nr-r^2)}{n(n-1)} \right) \\
&\le& e^{- \frac{2}{n(n-1)} \sum_{r=1}^n (nr-r^2)}
=e^{-(n+1)/3}.
\end{eqnarray*}
Hence, by Lemma \ref{lem:bound},
\begin{eqnarray*}
\left| P(W \in A) - \pi^m(A) \right| \le \frac{2.7314}{\sqrt{n}(n-1)}n(n-1)e^{-(n+1)/3}= 2.7314 \sqrt{n} e^{-(n+1)/3}.
\end{eqnarray*}
Taking supremum over $A$ and applying definition (\ref{def:dtv}) completes the proof.
\hfill $\Box$
\section{Bounds on the Stein equation}
\label{sec:bound}
In this section we present the proof of Lemma \ref{lem:bound}.
\noindent {\it Proof:} Let $m \in \{0,1\}$ be fixed. First, the equalities $f(m)=0$ and
\begin{eqnarray*}
f(x+2)=\frac{1_A(x)-\pi^m(A)+\beta_x f(x)}{\alpha_x} \quad \mbox{for $m<x \le n-2, x \in L_{m,n}$}
\end{eqnarray*}
specify $f(x)$ on $L_{m,n}$ uniquely, hence the solution to (\ref{Afr}) satisfying the given boundary condition is unique.
Next, with $r \in L_{m,n}$, we verify that $f^m_r(x)$ given by (\ref{frsol}) solves (\ref{Afr}) with $A=\{r\}$;
that $f^m_r(m)=0$ is given.
For $m<x<r,x \in L_{m,n}$, applying the balance equation (\ref{eqn:balance}) to obtain the second equality,
we have
\begin{eqnarray*}
\lefteqn{\alpha_x f^m_r(x+2)-\beta_x f^m_r(x) }\\
&=& \alpha_x \left( -\frac{\pi^m([0,x]\cap L_{m,n})\pi^m_r}{\beta_{x+2} \pi^m_{x+2}} \right) - \beta_x \left( - \frac{\pi^m([0,x-2]\cap L_{m,n})\pi^m_r}{\beta_x \pi^m_x} \right)\\
&=& \alpha_x \left( -\frac{\pi^m([0,x]\cap L_{m,n})\pi^m_r}{\alpha_x \pi^m_x} \right) - \beta_x \left( - \frac{\pi^m([0,x-2]\cap L_{m,n})\pi^m_r}{\beta_x \pi^m_x} \right)\\
&=&-\pi^m_r.
\end{eqnarray*}
If $x=r$ then
\begin{eqnarray*}
\lefteqn{\alpha_x f^m_r(x+2)-\beta_x f^m_r(x)}\\
&=& \alpha_r \left( \frac{\pi^m([r+2,n]\cap L_{m,n})\pi^m_r}{\beta_{r+2} \pi^m_{r+2}}\right)-\beta_r \left( \frac{-\pi^m([0,r-2]\cap L_{m,n})\pi^m_r}{\beta_r \pi^m_r}\right)\\
&=& \alpha_r \left(\frac{\pi^m([r+2,n]\cap L_{m,n})\pi^m_r}{\alpha_r \pi^m_r}\right)-\beta_r \left(\frac{-\pi^m([0,r-2]\cap L_{m,n})\pi^m_r}{\beta_r \pi^m_r}\right)\\
&=& \pi^m([r+2,n]\cap L_{m,n})+\pi^m([0,r-2]\cap L_{m,n})=1-\pi^m_r.
\end{eqnarray*}
If $x>r$ then
\begin{eqnarray*}
\lefteqn{\alpha_x f^m_r(x+2)-\beta_x f^m_r(x)}\\
&=& \alpha_x \left(\frac{\pi^m([x+2,n]\cap L_{m,n})\pi^m_r}{\beta_{x+2} \pi^m_{x+2}}\right)-\beta_x \left(\frac{\pi^m([x,n]\cap L_{m,n}) \pi^m_r}{\beta_x \pi^m_x}\right)\\
&=& \alpha_x \left(\frac{\pi^m([x+2,n]\cap L_{m,n})\pi^m_r}{\alpha_x \pi^m_x}\right)-\beta_x \left(\frac{\pi^m([x,n]\cap L_{m,n}) \pi^m_r}{\beta_x \pi^m_x}\right)\\
&=&-\pi^m_r.
\end{eqnarray*}
Hence $f^m_r(x)$ solves (\ref{Afr}).
Next, to consider the solution of (\ref{Afr}) more generally for $A \subset L_{m,n}$ and $x \in L_{m,n}$, letting
\begin{eqnarray*}
U_{m,x} = [0,x-2]\cap L_{m,n} \qmq{and} U_{m,x}^c = L_{m,n} \setminus U_{m,x},
\end{eqnarray*}
we may write (\ref{frsol}) more compactly as
\begin{eqnarray*}
f^m_r(x)=\frac{1}{\beta_x \pi^m_x}\left( \pi^m(U_{m,x}^c) \pi^m(\{r\} \cap U_{m,x}) - \pi^m(U_{m,x}) \pi^m(\{r\} \cap U_{m,x}^c) \right).
\end{eqnarray*}
By linearity, the solution of (\ref{Afr}) for $A \subset L_{m,n}$ is given by $f^m_A(m)=0$, and for $x>m,x \in L_{m,n}$,
by
\begin{eqnarray*}
f^m_A(x)=\frac{1}{\beta_x \pi^m_x}\left(\pi^m(U_{m,x}^c)\pi^m(A \cap U_{m,x})-\pi^m(U_{m,x})\pi^m(A \cap U_{m,x}^c) \right)
\end{eqnarray*}
(cf \cite{bhj}, p. 7), and so, for all $x \in L_{m,n}$,
\begin{eqnarray*}
-\frac{1}{\beta_x \pi^m_x}\pi^m(U_{m,x})\pi^m(U_{m,x}^c) \le f_A^m(x) \le \frac{1}{\beta_x \pi^m_x}\pi^m(U_{m,x}^c) \pi^m(U_{m,x}),
\end{eqnarray*}
or that
\begin{eqnarray}
\left| f^m_A(x) \right| \le \frac{1}{\beta_x \pi^m_x} \pi^m(U_{m,x}) \pi^m(U_{m,x}^c).\label{smalln}
\end{eqnarray}
Since $f^m_A(m)=0$ and the upper bound of Lemma~\ref{lem:bound} reduces to $\infty$ if $0 \le n\le 1$, we only need to bound $f^m_A(x)$ for $n\ge 2$ and $x\ge 2$. Direct computation using (\ref{smalln}) gives $|f_A^0(2)|\le 1/4$ for $n=2$, $|f_A^0(2)|\le 1/8$ and $|f_A^1(3)|\le 1/8$ for $n=3$, $|f_A^0(2)|=|f_A^0(4)|\le 7/96$ and $|f_A^1(3)|\le 1/12$ for $n=4$. Therefore, it remains to prove Lemma~\ref{lem:bound} for $n\ge 5$.
Noting that for $x\ge\frac n2+1$ we have $\beta_x\ge\left(\frac n2+1\right)\frac n2$, and for $x<\frac n2+1$ that $\alpha_{x-2}=(n-x+2)(n-x+1)>\left(\frac n2+1\right)\frac n2$,
using (\ref{eqn:balance}), we obtain from (\ref{smalln}) that
\begin{eqnarray} \label{bnd:by-fAx}
\left| f^m_A(x) \right| \le \left\{\begin{array}{ll}
\frac{\pi^m(U_{m,x}) \pi^m(U_{m,x}^c)}{\beta_x \pi^m_x} \le \frac{1}{\left( \frac{n}{2}+1 \right)\frac
{n}{2}}\frac{\pi^m(U_{m,x}) \pi^m(U_{m,x}^c)}{\pi^m_x}&\mbox{ if }x\ge \frac n2+1,\\
\frac{\pi^m(U_{m,x}) \pi^m(U_{m,x}^c)}{\alpha_{x-2} \pi^m_{x-2}}\le \frac{1}{\left( \frac{n}{2}+1 \right)\frac
{n}{2}}\frac{\pi^m(U_{m,x}) \pi^m(U_{m,x}^c)}{\pi^m_{x-2}}&\mbox{ if }x< \frac n2+1.
\end{array}\right.
\end{eqnarray}
Clearly, for $i \ge x$,
\begin{eqnarray*}
\frac{\pi^m_i}{\pi^m_x}=\frac{{n \choose i}}{{n \choose x}} =
\left\{
\begin{array}{cl}
1 & \mbox{if $i=x$}\\
\frac{(n-x) \cdots (n-i+1)}{(x+1) \cdots i} & \mbox{if $i \ge x+2$.}
\end{array}
\right.
\end{eqnarray*}
Hence, we can write, for $i \ge x+2$,
\begin{eqnarray} \label{prodforigexp2}
\frac{\pi^m_i}{\pi^m_x}= \left( \frac{n-x}{x+1}\right)\left( \frac{n-x-1}{x+2}\right) \cdots \left( \frac{n-i+1}{i} \right)=\prod_{y=0}^{i-x-1}\frac{n-x-y}{x+1+y}.
\end{eqnarray}
Note that as $(n-x)/(x+1) \le 1$ for $x \ge n/2$, the terms in the product (\ref{prodforigexp2}) are decreasing. In particular,
\begin{eqnarray} \label{star6}
\frac{\pi^m_i}{\pi^m_x} \le 1 \qmq{for $i \ge x$, and} \prod_{0 \le y \le
\lfloor \frac{\sqrt{n}}{2} \rfloor }\frac{n-x-y}{x+1+y} \le 1\, \mbox{ provided } x \ge \frac{n}{2}.
\end{eqnarray}
For $n$ even let $x_s=n/2$, and for $n$ odd let $x_s=(n-1)/2$ when $m=0$, and $x_s=(n+1)/2$ when $m=1$. Then, except for the case where $m=0$ and $x=(n+1)/2$, which we deal with separately, we have
$$
\pi^m(U_{m,x})\pi^m(U_{m,x}^c)=\pi^m(U_{m,2x_s-x+2})\pi^m(U_{m,2x_s-x+2}^c),
$$
and we may therefore assume $x \ge x_s+1$, and so $x \ge n/2+1$.
Since for $y \ge \sqrt{n}/2$, recalling $x \ge n/2+1$, we have
\begin{eqnarray}\label{star7}
\frac{n-x-y}{x+1+y} \le \frac{n-\left( \frac{n}{2}+1\right)-\frac{\sqrt{n}}{2}}{\left( 1+\frac{n}{2}\right)+1+\frac{\sqrt{n}}{2}}=\frac{\frac{n}{2}-\frac{\sqrt{n}}{2}-1}{\frac{n}{2}+2+\frac{\sqrt{n}}{2}}=1-\frac{\sqrt{n}+3}{\frac{n}{2}+2+\frac{\sqrt{n}}{2}},
\end{eqnarray}
applying (\ref{prodforigexp2}) and (\ref{star6}) we conclude that
\begin{eqnarray*}
\frac{\pi^m_i}{\pi^m_x} \le \left(1-\frac{\sqrt{n}+3}{\frac{n}{2}+2+\frac{\sqrt{n}}{2}} \right)^{i-x-\lfloor \frac{\sqrt{n}}{2} \rfloor-1} \quad \mbox{for $i \ge x+\lfloor \frac{\sqrt{n}}{2} \rfloor+1$}.
\end{eqnarray*}
Hence, applying (\ref{star6}) again, here to obtain the second inequality, we have
\begin{eqnarray}
\lefteqn{\frac{1}{\left( \frac{n}{2}+1\right) \frac{n}{2}} \frac{\pi^m(U_{m,x}^c)}{\pi^m_x}}\nonumber \\
&\le&
\frac{1}{\left(\frac{n}{2}+1 \right) \frac{n}{2}} \left( \sum_{x \le i \le x + \lfloor\frac{\sqrt{n}}{2}\rfloor,i\in L_{m,n}} \frac{\pi^m_i}{\pi^m_x} + \sum_{i \ge x + \lfloor\frac{\sqrt{n}}{2}\rfloor+1,i\in L_{m,n}} \frac{\pi^m_i}{\pi^m_x} \right) \nonumber \\
&\le& \frac{1}{\left( \frac{n}{2} + 1\right) \frac{n}{2}} \left( \left( \frac{\sqrt{n}}{4}+1\right) + \sum_{j=0}^\infty \left( 1- \frac{\sqrt{n}+3}{\frac{n}{2}+2+\frac{\sqrt{n}}{2}}\right)^{2j} \right)\nonumber \\
&=& \frac{1}{\left(\frac{n}{2}+1 \right) \frac{n}{2}} \left(\frac{\sqrt{n}}{4}+1+\frac{1}{1-\left(1-\frac{\sqrt{n}+3}{\frac{n}{2}+2+\frac{\sqrt{n}}{2}}\right)^2} \right)\nonumber\\
&\le& \frac{2.7314}{\sqrt{n}(n-1)} \quad \mbox{for $n \ge 1$.}\label{g1n}
\end{eqnarray}
This final inequality is obtained by determining the maximum of the function
$$
g_1(n):=\frac{1}{\left(\frac{n}{2}+1 \right) \frac{n}{2}} \left(\frac{\sqrt{n}}{4}+1+\frac{1}{1-\left(1-\frac{\sqrt{n}+3}{\frac{n}{2}+2+\frac{\sqrt{n}}{2}}\right)^2} \right)\sqrt{n}(n-1)
$$
by noting $g_1(n)<1+\frac4{\sqrt{n}}+\frac{(n+4+\sqrt{n})^2}{(n+2)(n+3\sqrt{n})}<2.5$ for $n\ge 64$ and $\max_{1\le n\le 63}g_1(n)=g_1(9)=2.7313131\ldots$.
Lastly we handle the situation where $n$ is odd, $m=0$ and $x=(n+1)/2=:x_0 \in L_{0,n}$, in which case $n=3\,{\rm mod}\,4$. In place of (\ref{star7}), we have, for $y \ge \sqrt{n}/2$,
\begin{eqnarray*}
\frac{n-x_0-y}{x_0+1+y} \le \frac{n-\left( \frac{n+1}{2}\right)-\frac{\sqrt{n}}{2}}{\left( \frac{n+1}{2}\right)+1+\frac{\sqrt{n}}{2}}=1-\frac{4+2\sqrt{n}}{n+3+\sqrt{n}}.
\end{eqnarray*}
Since (\ref{star6}) is valid for all $x \ge n/2$, in view of (\ref{prodforigexp2}) we obtain the bound
\begin{eqnarray*}
\frac{\pi^m_i}{\pi^m_{x_0}} \le \left(1-\frac{4+2\sqrt{n}}{n+3+\sqrt{n}} \right)^{i-x_0-\lfloor \frac{\sqrt{n}}{2} \rfloor-1} \quad \mbox{for $i \ge x_0+\lfloor \frac{\sqrt{n}}{2} \rfloor+1$}.
\end{eqnarray*}
Using (\ref{star6}) again for the first inequality we have
\begin{eqnarray}
\lefteqn{\frac{1}{\left( \frac{n}{2}+1\right) \frac{n}{2}} \frac{\pi^m(U_{m,x_0})\pi^m(U_{m,x_0}^c)}{\pi^m_{x_0}}}\nonumber \\
&=&
\frac{1}{2\left(\frac{n}{2}+1 \right) \frac{n}{2}} \left( \sum_{x_0 \le i \le x_0 + \lfloor\frac{\sqrt{n}}{2}\rfloor,i\in L_{m,n}} \frac{\pi^m_i}{\pi^m_{x_0}} + \sum_{i \ge x_0 + \lfloor\frac{\sqrt{n}}{2}\rfloor+1,i\in L_{m,n}} \frac{\pi^m_i}{\pi^m_{x_0}} \right) \nonumber \\
&\le& \frac{1}{\left( \frac{n}{2} + 1\right) n} \left( \left( \frac{\sqrt{n}}{4}+1\right) + \sum_{j=0}^\infty \left(1-\frac{4+2\sqrt{n}}{n+3+\sqrt{n}} \right)^j \right) \nonumber\\
&=& \frac{1}{\left(\frac{n}{2}+1 \right) n} \left(\frac{\sqrt{n}}{4}+1+\frac{n+3+\sqrt{n}}{4+2\sqrt{n}} \right)\nonumber\\
&\le& \frac{1.638496535}{\sqrt{n}(n-1)} \quad \mbox{for $n \ge 1$,}\label{g2n}
\end{eqnarray}
where the last inequality is from bounding the function
$$g_2(n):=\frac{1}{\left(\frac{n}{2}+1 \right) n} \left(\frac{\sqrt{n}}{4}+1+\frac{n+3+\sqrt{n}}{4+2\sqrt{n}}\right)(n-1)\sqrt{n},$$
with $g_2(n)\le \frac{1}{2}+\frac{2}{\sqrt{n}}+\frac{n+3+\sqrt{n}}{n+2\sqrt{n}}\le 1.6$ for $n\ge 400$ and $\max_{1\le n\le 399}g_2(n)=g_2(23)=1.638496535$.
The result now follows from combining the estimates (\ref{g1n}), (\ref{g2n}) and (\ref{bnd:by-fAx}).
\hfill $\Box$
We remark that a direct argument using Stirling's formula for the case $x=\lfloor n/2\rfloor$ shows that the best order that can be achieved for the estimate of $f^m_A$ is $O(n^{-3/2})$.
{\bf Acknowledgement:} The authors would like to thank the organizers of the conference held at the National University of Singapore in honor of Louis Chen's birthday for the opportunity to collaborate on the present work.
|
\section{Introduction and Summary}
According to the representation theory of the Poincar\'e group in 4D, a massive spin-2 state has to have
five degrees of freedom; these can be thought of as the helicity-$0$, $\pm 1$, $\pm 2$ states.
A good Lagrangian for the massive spin-2 has to be able to describe these states.
The Fierz-Pauli mass term \cite{FP} is the only ghost- and tachyon-free term at the quadratic
order that describes the above 5 states \cite{Nieu}. However, in the zero mass
limit it does not recover the linearized Einstein's gravity, since the helicity-$0$ mode
couples to the trace of the matter stress-tensor with strength equal to that of the helicity-2;
this is called the vDVZ discontinuity \cite{vDVZ}. If true, it would rule out massive
gravity on the grounds
of solar system observations. However, Vainshtein \cite{Arkady} argued that the troublesome
longitudinal mode is suppressed at measurable distances by nonlinear effects, making the nonlinear theory
compatible with current empirical data \cite{dvali}. On the other hand, in a broad class of models,
the same nonlinear terms that are responsible for the above-mentioned suppression, give rise to an
instability known as the Boulware-Deser (BD) ghost \cite{BD}. This ghost appears as a 6th degree of
freedom in the theory, and even though it is infinitely heavy on the Minkowski background, it becomes
sufficiently light on locally nontrivial backgrounds, thus invalidating the theory
\cite{nima,Creminelli, DeffayetRombouts,Andrei}.
More recently, however, using the effective field theory formalism of
\cite{nima}, it has been shown in Ref. \cite{general} that there exists
a two parameter family of nonlinear generalization of the linear
Fierz-Pauli theory, that is free of the BD ghost order-by order and
to all orders, at least in the decoupling limit.
Most importantly, it was shown in Ref. \cite{resum} that the absence of the
BD ghost in the decoupling limit is
such a powerful requirement that it leads to the resummation of the entire
infinite series of the terms in the effective Lagrangian. As a result,
a candidate theory of massive General Relativity free of BD ghost,
was proposed \cite{resum}.
Using the Hamiltonian analysis in the unitary gauge
it was shown that for a certain choice of the free parameters of the
theory, and in the 4th order in nonlinearities, the Hamiltonian constraint that
forbids the BD ghost is maintained in the theory of \cite{resum}. Note that the quartic order
is special, since the lapse necessarily enters nonlinearly in all massive theories
precisely in this order \cite {Creminelli}, and it may appear that the hamiltonian
constraint should necessarily be lost then. In spite of this, the constraint is maintained in a subtle
way for special theories, as was shown for a toy model in \cite {general},
and shown in the 4th order for massive GR in \cite{resum}.
The existence of the Hamiltonian constraint to all orders in the unitary gauge,
and for generic values of the parameters, was shown in Ref. \cite{rachel1},
using the method of dealing with the lapse and shift proposed in
\cite {general,resum}. Moreover, Ref. \cite{rachel1} has
also argued for a secondary constraint, that follows from the conservation of
the Hamiltonian constraint
\footnote{The argument of the existence of the secondary constraint was challenged
in \cite {Lham}.}. Very recently, the existence of the secondary constraint
was explicitly confirmed in Ref. \cite {Rachel_confirm}.
The absence of the BD ghost among the local fluctuations of
the theory of \cite {resum} in a generic gauge has been shown using the {St\"uckelberg}
decomposition \cite {Stu}, as well as the helicity decompositions \cite {Helic}
to quartic orders in nonlinearities (in the latter two
references, previous misconceptions in the literature claiming the presence of the BD
ghost were also clarified). Motivated by the above developments, in the present work
we will proceed to study certain subtle properties black holes (BH) in the
theory of \cite {resum}.
In the unitary gauge Lagrangian of the theory the object $h_{\mu \nu} \equiv g_{\mu \nu}- \eta_{\mu \nu}$,
is the gravitational analog of the Proca field of massive electrodynamics, describing all
the five modes of the graviton. The diffeomorphism invariance can be restored by
introducing the four scalar fields
$\phi^a$ (the {St\"uckelberg} fields) \cite{Siegel,nima,sergey}, and replacing the Minkowski metric by
the covariant tensor $\partial_\mu \phi^a \partial_\nu \phi^b \eta_{ab}$
\begin{eqnarray}
g_{\mu\nu}=\partial_\mu \phi^a \partial_{\nu} \phi^b \eta_{ab}+H_{\mu\nu},
\label{H1}
\end{eqnarray}
where $H_{\mu \nu}$ denotes the covariantized metric perturbation, and $\eta_{ab}={\rm diag} (-1,1,1,1)$.
The existence of the 4 {St\"uckelberg} scalars $\phi^a$ in this theory leads to the existence of new
invariants in addition to the ones usually encountered in GR (Ricci scalar, Ricci tensor square,
Riemann tensor square, etc); one new basic invariant is
$I^{ab}=g^{\mu\nu}\partial_\mu \phi^a \partial_\nu \phi^b$. Note that the unitary
gauge is set by the condition $\phi^a=x^\mu \delta_\mu^a$.
In this gauge, $I^{ab}=g^{\mu\nu}\delta_\mu^a \delta_\nu^b$. Hence, any inverse
metric that has divergence (even those which are innocuous in GR) would exhibit
a singularity in the invariant $I^{ab}$. Is this singularity of any significance?
The singularity in the above invariant does not necessarily affect the geodesic motion
of external observers -- the geodesic equation is identical to that of GR, and due to its
covariance, one could remove from it what would have been a coordinate
singularity in $g^{\mu\nu}$ in GR. However, one would expect the singularities in
$I^{ab}=g^{\mu\nu}\delta_\mu^a \delta_\nu^b$ to be a problem for fluctuations around
classical solutions exhibiting it. Since $g^{\mu\nu}$
could change signs on either side of the singularity, this could lead to emergence
of ghosts and/or tachyons in the fluctuations around a given classical solution.
In what follows, we will take a conservative point of view and will
only accept solutions that have non-singular $I^{ab}$. These arguments, in a somewhat
different form, have already been emphasized recently by Deffayet and Jacobson \cite {CedricJ}.
The above arguments give rise to the following seeming puzzle. On the one hand, according to
the Vainshtein mechanism \cite {Arkady}, spherically symmetric solutions of massive gravity
should approximate those of GR better and better, as we increase the mass of the source and
come closer to it. This would imply that the metric of a BH near its horizon should
very much be similar to that of GR. On the other hand, the conventional Schwarzschild metric -- if it
were the solution of massive gravity in unitary gauge -- would be singular according to
the arguments above.
We reiterate this central point in more general terms: In order for a metric to qualify as a valid
description of a BH configuration, the physical singularities must be absent at the horizon.
Then, in the unitary gauge of massive gravity the Schwarzschild-like metric
\begin{eqnarray}
ds^2=-(1-f)dt^2+\frac{dr^2}{1-f}+r^2d\Omega^2, \quad \text{with e.g.}\quad f=r_g/r,
\label{sh}
\end{eqnarray}
cannot be a legitimate BH solution of the theory.
The same applies to the metric of de Sitter (dS) space in the static coordinates for which $f=m^2r^2$.
Recently, interesting BH solutions of massive gravity have been found in
Refs. \cite{Koyama,theo,GruzinovMirbabayi} (for other interesting solutions,
which will not be discussed here, see, \cite{dato}- \cite{LuigiII}).
Following Koyama, Niz and Tasinato (KNT) \cite{Koyama}, one can start in the unitary
gauge, and consider a most general stationary spherically symmetric metric.
Then, using the method developed by KNT, very interesting full non-linear solutions for stars and black holes
with Minkowskian asymptotics were found by Gruzinov and Mirbabayi in
\cite{GruzinovMirbabayi}. These solutions do exhibit the Vainshtein mechanism, and therefore
are potentially viable classical solutions for stars and other compact objects
in massive gravity (although their stability still remains to be studied). Nevertheless,
it is not clear, as emphasized in \cite {GruzinovMirbabayi}, whether these are appropriate
solutions for BHs. Even in the best case solution, when all the GR invariants are finite,
the invariant $g^{\mu\nu}\partial_\mu \phi^a \partial_\nu \phi^b \eta_{ab}$ diverges,
\cite{GruzinovMirbabayi}. As noted above, this divergence does not affect the
geodesic motion of any external observer, however, we expect it to be a problem for
fluctuations.
Could there be any solution that avoids the above issue? The answer is positive, and the
resolution is in the identification of the unitary gauge to the
coordinate system in which the black hole has no horizon (the Kruskal-Szekeres, Eddington-Finkelstein,
or Gullstrand-Painlev\'e systems come to our mind). The most convenient one for our purposes
is the Gullstrand-Painlev\'e (GP) system, in which the metric has the following form
\begin{eqnarray}
ds^2=-dt^2+(d r\pm \sqrt{f}dt)^2+r^2d\Omega^2,
\label{pa}
\end{eqnarray}
and is free of horizon singularities. It corresponds to the frame of an in-falling observer
and covers half the whole space (for either choice of sign).
Furthermore, if one has the metric \eqref{pa} as a solution in unitary gauge, then the coordinate
transformation to the metric \eqref{sh} will lead us to a background with $\phi^a\neq x^a$. This
means that if the configuration is described by the metric \eqref{sh}, the presence of a
halo of helicity $\pm 1$ and/or $0$ fields around the BH
is unavoidable. We will show in the present work that massive gravity in unitary gauge
admits BH solutions precisely of this type.
Interestingly, the dS-Schwarzschild solution found in \cite {Koyama,KoyamaSA}
do happen to satisfy our conservative criterion of non-singularity.
However, the solution that we present here is not among the ones
of \cite {Koyama,KoyamaSA}.
One more point worth emphasizing is that the BH
solutions of \cite{GruzinovMirbabayi} do exhibit the ``helicity-0 hair''
(e.g., produce an extra scalar force), while the ones found in \cite {Koyama,KoyamaSA},
and in the present work do not. The status of the ``no-hair'' theorems in GR with
the galileon field (which should capture some properties of the helicity-0 of
massive gravity) will be discussed in Ref. \cite {Lam}.
A limitation of our work is that we only manage to find these exact analytic solutions
for a specific choice of the two free parameters of massive gravity.
Such a choice is peculiar since on the obtained background, as we will show,
the kinetic terms for both the vector and scalar fluctuations vanish in the decoupling limit.
This fact would imply infinitely strong interactions for these modes
(unless these modes happen to be nondynamical
to all orders, e.g., due to the specific choice of the coefficients of the theory).
Because of this issue, we would like to regard the solutions obtained here as just
examples demonstrating how non-singular solutions should emerge. We also hope that
our solutions are representative of a broader class of solutions which
may have better behaved fluctuations.
In this regard, there seems to be a few directions in which the studies of massive gravity
BH's can be extended. First, one could look at the metric in the
unitary gauge which would be some generalization
of the Kruskal-Szekeres form. Second, one can extend the massive theory
of \cite {resum} by adding more degrees of freedom
to the existing 5 helicity states of massive graviton. In fact,
two consistent extensions have already been discussed so far: (I) adding one real
scalar field that makes the graviton
mass dynamical \cite {nonFRW}; (II) adding one massless tensor field with two degrees
of freedom \cite {HRbi} that makes the internal space metric of the {St\"uckelberg} field dynamical (bigravity).
In the latter case cosmological solutions were found
recently in \cite {VolkovII} and \cite {LuigiII}, while BH's
were studied in \cite {Pilo}.
\vspace{0.1in}
The work is organized as follows. Section 2 gives a brief review of the
theory of massive gravity \cite {resum}. In section 3 we find an
exact Schwarzschild-de Sitter solution, and in section 4 an
exact Reissner-Nordstr\"om-de Sitter, solution which have nonsingular $I^{ab}$.
These solution are similar to those discovered by Koyama, Niz and Tasinato, and
by Th.~Nieuwenhuizen, but differ in detail: our dS solution has no ghost
even though the vector field is present (compare to \cite {KoyamaSA}).
Moreover, on the obtained solution the
singularities in the invariant $I^{ab}$ are absent (compare to \cite {theo}).
In the Appendix A we give another exact Schwarzschild solution that
asymptotes to a conformaly rescaled Minkowski space, and briefly mention its peculiarities.
In the Appendix B we discuss fluctuations on the selfaccelerated solution of section 3.
\section{The Theory}
A massive graviton is described by the Lagrangian density of \cite {resum} specified below
\begin{eqnarray}
\mathcal{L}=\frac{M_{\rm pl}^2}{2}\sqrt{-g}\left(R + m^2 \mathcal U(g,\phi^a)\right),
\end{eqnarray}
where $\mathcal{U}$ is the potential for the graviton that depends on two free
parameters $\alpha_{3,4}$
\begin{eqnarray}
\mathcal U(g,\phi^a)=\(\mathcal{U}_2+\alpha_3\ \mathcal{U}_3+\alpha_4\ \mathcal{U}_4 \),
\label{eq:fullU}
\end{eqnarray}
and the individual terms in the potential are defined as follows:
\begin{eqnarray}
\label{po}
\mathcal{U}_2&=&[\mathcal{K}]^2-[\mathcal{K}^2]\,,\\
\mathcal{U}_3&=&[\mathcal{K}]^3-3 [\mathcal{K}][\mathcal{K}^2]+2[\mathcal{K}^3]\,,\\
\mathcal{U}_4&=&[\mathcal{K}]^4-6[\mathcal{K}^2][\mathcal{K}]^2+8[\mathcal{K}^3]
[\mathcal{K}]+3[\mathcal{K}^2]^2-6[\mathcal{K}^4]\,,
\label{pot}
\end{eqnarray}
where $\mathcal{K}^\mu_\nu(g,\phi^a)=\delta^\mu_\nu-\sqrt{g^{\mu\alpha}\partial_\alpha \phi^a
\partial_\nu \phi^b \eta_{ab}}$; rectangular brackets denote
traces, $[\mathcal{K}]\equiv {\rm Tr} (\mathcal{K})= \mathcal{K}^\mu_\mu$. The above potential is unique --
no further polynomial terms can be added to the action
without introducing the BD ghost.
The tensor $H_{\mu\nu}$ represents the covariantized metric perturbation, as discussed in the
introduction, which reduces to the $h_{\mu \nu}$ in unitary gauge.
While in a gauge unfixed theory we have
\begin{eqnarray}
H_{\mu \nu}=g_{\mu \nu}-\partial_\mu \phi^a \partial_\nu \phi^b \eta_{ab}.
\end{eqnarray}
Moreover, $\mathcal{U}$ is constructed in such a way that the theory admits the Minkowski background
\begin{eqnarray}
g_{\mu\nu}=\eta_{\mu\nu},\qquad \phi^a=x^\mu\delta_\mu^a.
\end{eqnarray}
Hence, it is natural to split $\phi$'s as the background plus the `pion' contribution
$\phi^a=x^a-\pi^a,$ and as it
was already mentioned in the introduction, the unitary gauge is defined by the condition $\pi^a=0$.
In the non-unitary gauge, on the other hand, it proves to be useful to adopt the following decomposition
\begin{eqnarray}
\pi^a=\frac{mA^a+\partial^a\pi}{\Lambda^3},
\label{decomp}
\end{eqnarray}
where $A^\mu$ describes in the decoupling limit the helicity $\pm 1$, while $\pi$ is the longitudinal
mode of the graviton (in the decoupling limit \cite{nima},
$M_{\rm pl}\rightarrow \infty$ and $m\rightarrow 0$, while $\Lambda^3\equivM_{\rm pl} m^2$ is held fixed).
This limit captures the approximation in which the energy scale is much greater than
the graviton mass scale, $E\gg m$.
For convenience, in what follows, we define the coefficients $\alpha$ and $\beta$ which are related
to those of \eqref{eq:fullU} by $\alpha_3=-(-\alpha+1)/3$ and $\alpha_4=-\beta/2+(-\alpha+1)/12$.
For generic values of the parameters $\alpha$ and $\beta$ the theory exhibits
the Vainshtein mechanism, as show in the decoupling limit
\cite {general}, and beyond \cite {Koyama,DatoGiga}. As was emphasized in \cite {general},
for one special choice, $\alpha =\beta=0$, the nonlinear interactions vanish
in the decoupling limit with fixed $\Lambda$, leaving the theory weakly coupled
(i.e., no Vainshtein mechanism) in this limit.
For this particular choice of the coefficients the action of massive gravity with the potential
\eqref{po}-\eqref{pot} (which can be rewritten in terms of just $[K]$ and tuned to it cosmological constant
\cite {Rachel0}, referred as a minimal model in Ref. \cite {Rachel0}), was shown not
to exhibit the Vainshtein mechanism also away from the decoupling limit \cite {Koyama}.
\section{A Black Hole on de Sitter}
In this section we present the Schwarzschild-de Sitter solution in the theory of
massive gravity described above. The obtained solution is free
of singularities (except from the conventional one appearing in GR).
For convenience we choose unitary gauge for the metric. In this gauge
the symmetric tensor $g_{\mu\nu}$ is an observable describing all the five degrees of
freedom of a massive graviton. The equations of motion in empty space read as follows
\begin{eqnarray}
G_{\mu\nu}+m^2 X_{\mu\nu}=0\,,
\label{ein}
\end{eqnarray}
where $X_{\mu\nu}$ is the effective energy-momentum tensor due to the graviton mass,
\begin{eqnarray}
&&X_{\mu \nu}=-\frac12\Big[
\mathcal{K}g_{\mu \nu}-{\cal K}_{\mu \nu}+\alpha\({\cal K}^2_{\mu \nu}-{\cal K} {\cal K}_{\mu \nu}+\frac 12 g_{\mu \nu} \([{\cal K}]^2-[{\cal K}^2]\)\)\\
&&+6 \beta\({\cal K}_{\mu \nu}^3-{\cal K} {\cal K}_{\mu \nu}^2+\frac 12 {\cal K}_{\mu \nu} \([{\cal K}]^2-[{\cal K}^2]\)-\frac 16 g_{\mu \nu} \([{\cal K}]^3-3 [{\cal K}][{\cal K}^2]+2[{\cal K}^3]\)\)
\Big]\notag\,.
\end{eqnarray}
Using the Bianchi identities, from \eqref{ein} we obtain
the following constraint on the metric
\begin{eqnarray}
m^2 \nabla^{\mu}X_{\mu\nu}=0,
\label{const}
\end{eqnarray}
where $\nabla ^\mu$ denotes the covariant derivative.
In order to obtain the expression for $X_{\mu \nu}$ we make use of the fact that the
Lagrangian is written as the trace of the polynomial of the matrix $\mathcal{K}^\mu_\nu$.
Thus, following the method by Koyama, Niz and Tasinato \cite {Koyama}, we choose the basis which diagonalizes the expression appearing
under the square root in the definition of $\mathcal{K}^\mu_\nu$ [One should bear in mind that this is
not a coordinate transformation, but rather a trick to simplify the procedure of getting the
equations of motion]. As a result, the potential becomes a function of the components of the
inverse metric, rather than the combination of square roots of matrices. Having done this,
one is free to vary the action with respect to the inverse metric components to obtain explicit
expression for \eqref{ein}. Since these expressions are quite cumbersome we will not
give them here.
Below, we concentrate on one particular family of the ghost-free theory of
massive gravity in which there is the following relations between the two free coefficients:
\begin{eqnarray}
\beta=-\frac{\alpha^2}{6}\,.
\label{fam1}
\end{eqnarray}
That this choice of the coefficients is special was first shown by Th. Nieuwenhuizen \cite {theo} (see also
\cite{GruzinovMirbabayi}). In particular, it was shown in \cite {theo} that for this choice the equation
\eqref{const} is automatically satisfied for a certain diagonal (in spherical coordinates)
and time-independent metrics. It is interesting, however, that the above property
persists for a more general class of non-diagonal spherically-symmetric metrics written as follows:
\begin{eqnarray}
ds^2=-A(r)dt^2+2B(r)dtdr+C(r)dr^2+w^2r^2d\Omega^2,
\label{anz}
\end{eqnarray}
where $w$ is a constant, while $A(r)$, $B(r)$ and $C(r)$ are arbitrary functions.
In subsection 3.1 we find an exact de Sitter solution to \eqref{ein},
and in subsection 3.2 we find an exact BH solution on the obtained
dS background. Note that the dS background is entirely due to the graviton mass.
\subsection{The de Sitter Solution}
We note that we would find an exact dS solution if we
required that
\begin{eqnarray}
m^2 X_{\mu \nu}=\lambda g_{\mu \nu},
\label{lambdacc}
\end{eqnarray}
where $\lambda$ is some constant. The solution of the equations \eqref{ein}
that also satisfies (\ref {lambdacc}) with a positive but otherwise arbitrary
$\alpha$ is given by
\begin{eqnarray}
ds^2=-\kappa ^2 dt^2+\left(\frac{\alpha}{\alpha+1}dr\pm\kappa{\sqrt{\frac{2}{3\alpha}}
\frac{\alpha}{(\alpha+1)}mr} dt\right)^2+\frac{\alpha^2}{(\alpha+1)^2}r^2d\Omega^2.
\label{pain}
\end{eqnarray}
Here, $\kappa$ is a positive integration constant. It is straightforward to check that
for \eqref{pain} we have $X_{\mu\nu}=(2/\alpha) ~g_{\mu\nu}$, leading to the expression for the Ricci scalar
\begin{eqnarray}
R=\frac{8}{\alpha}m^2,
\end{eqnarray}
as expected. Hence, this is a dS space with curvature scale
set by the graviton mass and one free parameter $\alpha$. One could imagine that
$m\sim (0.1- 1) H_0$, and $\alpha \sim (0.01 -1)$, in which case
the obtained dS solution (if stable) could describe dark energy.
Up to a rescaling of the coordinates,
the expression \eqref{pain} looks exactly like the de Sitter solution of GR
written in the Gullstrand--Painlev\'e frame. Either $\pm$ solution covers half of dS space.
One can rotate the obtained
solution to the static coordinate system at the expense of
nonzero {St\"uckelberg} fields. This will be done in the next subsection.
In either form, the solution has no additional singularities.
\subsection{Schwarzschild-de Sitter Background}
Having the solution of the previous subsection worked out, it is
straightforward to show that the system of equations \eqref{ein}
admits the following exact solution
\begin{eqnarray}
ds^2=-\kappa ^2 dt^2+\left({\tilde \alpha} dr\pm\kappa\sqrt{\frac{r_g}{{\tilde \alpha} r}+
\frac{2{\tilde \alpha}^2 }{3\alpha} m^2r^2} dt\right)^2+ {\tilde \alpha}^2 r^2d\Omega^2\,,
\label{pain1}
\end{eqnarray}
where ${\tilde \alpha } \equiv \alpha/(\alpha+1)$, and as before,
$\kappa$ is an integration constant. In order to bring this solution to
a more familiar form let us perform the following rescaling
\begin{eqnarray}
r&\rightarrow& \frac{\alpha+1}{\alpha} r,\nonumber\\
dt&\rightarrow& \frac{1}{\kappa}dt.
\label{trans}
\end{eqnarray}
The resulting metric reads
\begin{eqnarray}
ds^2=-dt^2+\left(dr\pm\sqrt{\frac{r_g}{r}+\frac{2}{3\alpha}m^2r^2} dt\right)^2+r^2d\Omega^2.
\label{pain2}
\end{eqnarray}
This is the Schwarzschild-de Sitter solution in the GP coordinates.
However, the above rescaling takes us away from the unitary gauge
\begin{eqnarray}
&&\phi^0=t\rightarrow t-\left(1- \frac{1}{\kappa} \right) t,\\
&&\phi^r=r\rightarrow r+\frac{1}{\alpha} r.
\end{eqnarray}
In terms of the `pions', $\pi^\mu\equiv x^\mu-\phi^\mu$, which can be decomposed as
$\pi^\mu=(mA^\mu+\partial^\mu\pi)/\Lambda^3$, we have
\begin{eqnarray}
\pi&=&\frac{\Lambda^3}{2}\left[-\left( 1-\frac{1}{\kappa} \right)t^2-\frac{1}{\alpha}r^2\right],\\
A^\mu&=&0.
\label{asig}
\end{eqnarray}
The fields in (\ref {asig}) correspond to the canonically normalized fields carrying
the helicity eigenstates in the decoupling limit.
\vspace{0.1in}
Let us now rewrite our solution into a more familiar coordinate system.
The metric can be transformed to a static slicing by means of the following
coordinate transformation
\begin{eqnarray}
dt\rightarrow dt+f'(r)dr,
\label{trans1}
\end{eqnarray}
with $f'(r)\equiv -g_{01}/g_{00}$ given by
\begin{eqnarray}
f'(r)=\pm \frac{\sqrt{\frac{r_g}{r}+\frac{2}{3\alpha}m^2r^2}}{1-\frac{r_g}{r}-\frac{2}{3\alpha}m^2r^2}.
\end{eqnarray}
The resulting expression for the metric reads as follows:
\begin{eqnarray}
ds^2=-\left(1-\frac{r_g}{r}-\frac{2}{3\alpha}m^2r^2\right)dt^2+\frac{dr^2}
{1-\frac{r_g}{r}-\frac{2}{3\alpha}m^2r^2}+r^2d\Omega^2.
\label{static}
\end{eqnarray}
This is nothing but the metric of the Schwarzschild-de Sitter
solution of GR in the static coordinates. However, this metric should be accompanied
by a nontrivial backgrounds for the {St\"uckelberg} fields. Indeed, it is evident that \eqref{trans1}
gives rise to the shift $\delta \phi^0=f(r)$. In turn, this gives rise to
a background for the `vector mode'
\begin{eqnarray}
A^0&=&-\frac{\Lambda^3}{\kappa m}f(r),\\
A^i&=&0.
\label{asig0}
\end{eqnarray}
This particular field assignment has been chosen according to scaling in
the decoupling limit. Namely, $f(r)$ vanishes in the decoupling limit linearly in $m$
hence it was ascribed to the ``vector mode''.
\vspace{0.1in}
We would like to make two important comments in the remainder of this section.
The first one concerns the integration constant $\kappa$.
Although, all the invariants of GR are independent of $\kappa$,
the new invariant that is characteristic of massive gravity
\begin{eqnarray}
I^{ab}\equiv g^{\mu\nu}\partial_\mu \phi^a \partial_\nu \phi^b,
\label{newinv}
\end{eqnarray}
does depend on this integration constant; in the unitary gauge $I^{ab}$ is just
the inverse of the GP metric \eqref{pain} which reads as follows
\begin{eqnarray}
\footnotesize{\left(
\begin{array}{ccc} \nonumber
-\frac{1}{\kappa^2} & \pm \frac{1}{\kappa} \frac{\alpha+1}{\alpha}\sqrt{\frac{\alpha+1}{\alpha}\frac{r_g}{r}+\frac{2}{3\alpha}\frac{\alpha^2}{(\alpha+1)^2}m^2r^2} & 0 \\
\pm \frac{1}{\kappa} \frac{\alpha+1}{\alpha}\sqrt{\frac{\alpha+1}{\alpha}\frac{r_g}{r}+\frac{2}{3\alpha}\frac{\alpha^2}{(\alpha+1)^2}m^2r^2} & \left( \frac{\alpha+1}{\alpha} \right)^2\left( 1-\frac{\alpha+1}{\alpha}\frac{r_g}{r}-\frac{2}{3\alpha}\frac{\alpha^2}{(\alpha+1)^2}m^2r^2 \right) & 0 \\
0 & 0 & \left( \frac{\alpha+1}{\alpha} \right)^2 \Omega^{-1}_{2\times 2}
\end{array}
\right)}.
\end{eqnarray}
Thus, the backgrounds with different values of $\kappa$ correspond to distinct super-selection sectors
labeled by the values of $I^{ab}$.
The second comment concerns the issue of small fluctuations on top of this solution.
One may worry that the scalar perturbations on this background are infinitely strongly
coupled in the light of the results of \cite{dato}. In the latter work it was found that,
for the parameters chosen as in \eqref{fam1}, the de Sitter background has
infinitely strongly coupled fluctuations in the decoupling limit. However, we should point out that
the self-accelerated background discussed in this section is different from that
studied in \cite{dato}. This distinction is manifest in \eqref{asig0} by the presence of the
background for $A_0$, which vanishes in the case of \cite{dato}.
Still, one could argue that it is unnecessary to perform the transformation of
variables \eqref{trans1} responsible for this difference, and limit oneself
to the rescaling of the coordinates
\begin{eqnarray}
r \rightarrow \frac{\alpha+1}{\alpha} r\,,~~~t \rightarrow \frac{1}{\kappa}t\,,
\label{hyp}
\end{eqnarray}
which clearly does not give rise to the vector background.
As a result, the `pion' configuration will become similar to that of \cite{dato}, while the metric itself
will be quite different, namely\footnote{For simplicity we set $r_g=0$ and drop the numerical factors.}
\begin{eqnarray}
ds^2=-(1-m^2r^2)dt^2+2mrdtdr+dr^2+r^2d\Omega^2.
\label{met}
\end{eqnarray}
Now, if we were to take this metric as the one in which the decoupling limit should be taken, then
we would find that the gauge freedom that is left in this limit
\begin{eqnarray}
g_{\mu \nu} \rightarrow g_{\mu \nu}+\partial _{( \mu} \xi _{\nu )},
\end{eqnarray}
would not be enough for bringing \eqref{met} to the form of de Sitter space in either
conformal or static slicing. Furthermore, the canonically normalized \eqref{met} diverges in the decoupling limit, in such a way that this divergence can be isolated only in the vector mode. If so, then
no conclusion can be drawn about the perturbations around our background based on the
results of \cite{dato}. This, on the other hand, does not necessarily imply that the
fluctuations are fine. As we show in the Appendix B the vector and scalar
fluctuations may be infinitely strongly coupled.
\subsection{From Gullstrand-Painlev\'e to Kruskal-Szekeres}
In this section we ask the question whether the BH solution of the GP form could be analytically continued to cover the other half of the space-time as well. This can be done by going to the Kruskal-Szekeres (KS) coordinates and analyzing the {St\"uckelberg} fields.
Let us start addressing this point by considering the following background
\begin{eqnarray}
ds^2=-dt_{GP}^2+\left(dr+\sqrt{\frac{r_g}{r}}dt_{GP}\right)^2+r^2 d\Omega^2, \qquad \phi^a=x^a.
\label{metric}
\end{eqnarray}
First we rewrite the metric in static slicing by performing the following change of the time variable
\begin{eqnarray}
\phi^0=t_{GP}=t+2r_g \sqrt{\frac{r}{r_g}}+r_g \text{ ln}\left( \frac{\sqrt{\frac{r}{r_g}}-1}{\sqrt{\frac{r}{r_g}}+1} \right).
\end{eqnarray}
As a result the metric takes on the Schwarzschild form.
In order to go to KS coordinates we use reparametrizations identical to the one used in GR
\begin{eqnarray}
&&\left(\frac{r}{r_g}-1\right)e^{r/r_g}=X^2-T^2, \\
&&t=r_g\text{ ln}\left( \frac{X+T}{X-T} \right).
\end{eqnarray}
For the analysis of the $\phi$'s in KS coordinates we make the `near the horizon' approximation $r/r_g\rightarrow 1$, since this is the the region of our interest. In this limit the above coordinate transformation simplifies to (near the horizons $T=\pm X$, with signs corresponding to the black- and white-hole respectively)
\begin{eqnarray}
&&\left(\frac{r}{r_g}-1\right)=\frac{1}{e}(X^2-T^2), \\
&&t=r_g\text{ ln}\left( \frac{X+T}{X-T} \right).
\end{eqnarray}
As a result the $\phi$'s take the following form (using the fact that $T^2-X^2$ is small)
\begin{eqnarray}
&&\phi^0=2r_g\text{ ln}(X+T)+r_g (\text{ln}(1/4)+1),\\
&&\phi^r=r_g\left(1+\frac{1}{e}(X^2-T^2)\right).
\end{eqnarray}
Notice that $\phi^0$ is singular on the horizon of the white hole (while being regular on the black hole horizon).
The metric in these coordinates is given by
\begin{eqnarray}
ds^2=4r_g^3\frac{e^{-r/r_g}}{r}(-dT^2+dX^2)+r^2d\Omega^2
\end{eqnarray}
The invariant $I^{ab}=g^{\mu\nu}\partial_\mu \phi^a \partial_\nu \phi^b$ on the above background is singular at $X=-T$, corresponding to the horizon of the white hole.
If one takes the original GP metric \eqref{metric} to describe the white hole instead of the black hole (this is achieved by flipping the relative sign of the expressions in parentheses) then after analytical continuation to KS coordinates the singularity will appear on the black hole horizon rather than on the one of the white hole. The generalization of this arguments for the case of the dS is straightforward.
\section{Reissner-Nordstr\"om solution on de Sitter}
The ghost-free theory of massive gravity with $\beta=-\alpha^2/6$, upon its coupling to the Maxwell's theory of electromagnetism, possesses the following Reissner-Nordstr\"om solution on dS space
\begin{eqnarray}
ds^2=-dt^2+\left(\tilde{\alpha}dr\pm\sqrt{\frac{r_g}{\tilde{\alpha}r}+\frac{2\tilde{\alpha}^2}{3\alpha} m^2r^2-\frac{\tilde{Q}^2}{\tilde{\alpha}^4 r^2}} dt\right)^2+\tilde{\alpha}^2 r^2d\Omega^2,
\label{GPcharged}
\end{eqnarray}
with $\tilde{\alpha} \equiv \alpha/(\alpha+1)$ and the electromagnetic field given by
\begin{eqnarray}
E=\frac{\tilde{Q}}{r^2} \qquad \text{and} \qquad B=0.
\end{eqnarray}
In order to normalize the radial coordinate appropriately and to rewrite the solution in the static slicing, one needs to perform the rescaling
\begin{eqnarray}
r\rightarrow \frac{r}{\tilde{\alpha}},
\end{eqnarray}
supplemented with the following transformation of time
\begin{eqnarray}
dt\rightarrow dt+f'(r)dr, \quad \text{with} \quad f'(r)\equiv -\frac{g_{01}}{g_{00}}=\pm \frac{\sqrt{\frac{r_g}{r}+\frac{2}{3\alpha} m^2r^2-\frac{\tilde{Q}^2}{\tilde{\alpha}^2 r^2}}}{1-\frac{r_g}{r}-\frac{2}{3\alpha} m^2r^2+\frac{\tilde{Q}^2}{\tilde{\alpha}^2 r^2}}.
\end{eqnarray}
As a result the metric takes the familiar form
\begin{eqnarray}
ds^2=-(1-\frac{r_g}{r}-\frac{2}{3\alpha} m^2r^2+\frac{\tilde{Q}^2}{\tilde{\alpha}^2 r^2})dt^2+\frac{dr^2}{1-\frac{r_g}{r}-\frac{2}{3\alpha} m^2r^2+\frac{\tilde{Q}^2}{\tilde{\alpha}^2 r^2}}+r^2d\Omega^2,
\end{eqnarray}
while the {St\"uckelberg} fields become
\begin{eqnarray}
\phi^0&=&t+f(r), \\
\phi^r&=&r+\frac{1}{\alpha} r.
\end{eqnarray}
In this reference frame the electromagnetic field is
\begin{eqnarray}
E=\frac{\tilde{Q}}{\tilde{\alpha} r^2} \qquad \text{and} \qquad B=0.
\end{eqnarray}
And, for obvious reasons the actual charge should be defined by $Q\equiv \tilde{Q}/\tilde{\alpha}$.
\vskip 10pt
\subsection*{Acknowledgments}
We'd like to thank Gia Dvali, Lam Hui, Mehrdad Mirbabayi, Alberto Nicolis,
and David Pirtskhalava for useful comments. The work of LB and GC are supported by
the NYU James Arthur and MacCraken Fellowships, respectively. CdR is supported by the
Swiss National Science Foundation. GG is supported by NSF grant PHY-0758032. AJT would like to thank the
Universit\'e de Gen\`eve for hospitality whilst this work was being completed.
\renewcommand{\theequation}{\Roman{equation}}
\setcounter{equation}{0}
|
\section{Introduction}
Accreting X-ray pulsars, mostly residing in high-mass X-ray binaries (HMXBs),
often show abrupt increases in their X-ray luminosity lasting from
a few tens of seconds to several hours -- X-ray flares. Flaring activity
is often observed on top of a slower flux variation related to
X-ray outbursts or super-orbital modulation. Among the best-known
``flaring'' X-ray pulsars are LMC\,X-4 \citep[e.g.][end references
therein]{Moon:etal:03}, SMC\,X-1 \citep{Angelini:etal:91,Moon:etal:03b},
Vela\,X-1 \citep{Kreykenbohm:etal:08,Fuerst:etal:10}. X-ray binaries with
a donor star of Be (or Oe) spectral type (Be/X-ray binaries or BeXRB)
are currently the most numerous class of HMXBs with X-ray
pulsars \citep{Liu:etal:06}, although another class of HMXB pulsars,
the supergiant fast X-ray transients (SFXT), is rapidly growing
\citep[e.g.][]{Sidoli:11}. BeXRBs are characterized by periodic or sporadic
X-ray outbursts lasting from several days to several weeks when the
neutron star accretes matter from the equatorial disk around the
donor star \citep[see e.g.][for a recent review]{Reig:11}.
These systems are also known
to show occasional X-ray flares \citep{Finger:etal:99,Reig:etal:08,Caballero:etal:08,
Postnov:etal:08}.
Probably the most remarkable flaring activity among BeXRBs was exhibited
by the 42\,s pulsar \object{EXO\,2030+375}.
A series of six roughly equidistant flares with a mean
period of $\sim$four hours was observed with the EXOSAT satellite
a few months after the giant outburst of the source in 1985
\citep{Parmar:etal:89a}.
EXO\,2030+375 is one of the most regularly monitored BeXRBs.
In addition to
two ``major'' outbursts in 1985 and 2006 with a peak X-ray luminosity
of $L_{\rm X}\gtrsim 10^{38}$\,erg\,s$^{-1}$, assuming a distance of 7\,kpc
\citep{Wilson:etal:02}, the source exhibited less
powerful ``normal'' outbursts with
$L_{\rm X}\sim 10^{37}$\,erg\,s$^{-1}$ almost every orbit. The orbital
period of the system is $\sim$46\,d \citep{Wilson:etal:08}.
Some flaring of EXO\,2030+375 has been reported since the
EXOSAT observations, but it has generally appeared to be a single
flare during the rise of normal outbursts
\citep{Reig:Coe:98,Camero:etal:05}. Flares were also apparently present
during and between the normal outbursts shortly after the 2006 major outburst
\citep[Fig. 3 in][]{Wilson:etal:08}. But the short duration and
broad spacing of the observations make it unclear whether this is the same
quasi-periodic phenomenon that was seen with EXOSAT.
In this Letter we present the INTEGRAL observations of the source
that, for the first time since the EXOSAT observations, reveal
strong quasi-periodic flaring behavior.
\section{Observations and data processing}
Since its major outburst in 2006, EXO\,2030+375 has repeatedly appeared in the
field of view of \textit{International Gamma-Ray Astrophysics Laboratory}
(INTEGRAL, \citealt{Winkler:etal:03}) during the
observational program concentrated on the ``Cygnus Region''.
The INTEGRAL observatory has three main X-ray
instruments: (i) the imager IBIS sensitive from $\sim$20\,keV to a
few MeV with high spatial and moderate spectral resolution
\citep{Ubertini:etal:03}; (ii) the spectrometer SPI, which is
sensitive in the roughly
same energy range as IBIS, but with much higher spectral resolution and
substantially lower imaging capability \citep{Vedrenne:etal:03};
and (iii) the X-ray monitor JEM-X with moderate spectral and spatial
resolution, operating between $\sim$3 and $\sim$35\,keV \citep{Lund:etal:03}.
INTEGRAL observations normally consist of a series of pointings called
Science Windows, 2 to 4\,ksec each.
\begin{figure}
\centering
\resizebox{\hsize}{!}{\includegraphics{lc_obs.eps}}
\caption{INTEGRAL observations of EXO\,2030+375 (vertical bars)
superposed on the Swift/BAT
light curve of the source since its major outburst in
June--October 2006.
}
\label{fig:lcobs}
\end{figure}
Because the times of the observations were not optimized for EXO\,2030+375,
the INTEGRAL coverage of the source's normal outbursts is rather sparse.
The Swift/BAT light curve of the pulsar\footnote{We used the Swift/BAT
transient monitor results provided by the Swift/BAT team}
with the indicated INTEGRAL observations is shown in Fig.\,\ref{fig:lcobs}.
For our analysis we used the ISGRI detector layer of IBIS sensitive in the
20--300\,keV energy range \citep{Lebrun:etal:03} and JEM-X.
Owing to limited count-rate statistics, no additional information
could be gained from SPI data.
The standard data processing was performed with
version 9 of the Offline Science Analysis (OSA) software
provided by the INTEGRAL Science Data Centre
(ISDC, \citealt{Courvoisier:etal:03}). We performed an
additional gain correction of the ISGRI energy scale based
on the background Tungsten spectral lines.
To search for the flaring activity, we examined the entire IBIS/ISGRI
light curve by combining the publicly available ISGRI data products
in the HEAVENS data base\footnote{http://www.isdc.unige.ch/heavens/heavens}
with the results of our own analysis. We did not find any clear
evidence of flares in all the data except for the latest
INTEGRAL observations of the source
in November and December 2010 (MJD $\sim$55520--55540),
partially covering a normal outburst.
The rising part of the outburst is shown in Fig.\,\ref{fig:lcflares}.
The upper panel of Fig.\,\ref{fig:nv} shows the entire
outburst as observed with INTEGRAL.
In this work we concentrate on the analysis of the source's flaring
behavior during this outburst.
\begin{figure}
\resizebox{0.9\hsize}{!}{\includegraphics{lcflares.eps}}
\caption{The INTEGRAL/ISGRI light curve of EXO\,2030+375 during the
rise of its normal outburst in November--December 2010.
The flaring activity is clearly seen.
The light curve of the entire
outburst is shown in Fig.\,\ref{fig:nv}.
One Crab corresponds to $\sim$260 cts/s in the specified energy range.}
\label{fig:lcflares}
\end{figure}
\section{Timing analysis}
As can be seen in Fig.\,\ref{fig:lcflares}, the flux in
the rising part of the outburst experiences quasi-periodic
oscillations/flares that cease as the averaged flux increases.
One can identify at least five subsequent flares with a
mean period of $\sim$0.3 days ($\sim$7 hours).
\begin{figure}
\resizebox{0.95\hsize}{!}{\includegraphics{nv.eps}}
\caption{\emph{Top}: ISGRI light curve of the EXO\,2030+375
outburst approximated
with a polynomial function (solid curve) which represents
the averaged evolution of the flux.
\emph{Bottom:} Normalized excess variance of the source flux with respect to the
polynomial function.
}
\label{fig:nv}
\end{figure}
To characterize the level of the flux variability during the outburst we
calculated the normalized excess variance in the light curve in relatively
broad time intervals.
We defined four equal adjacent intervals in the rising part and five
equal adjacent intervals in the decay.
The normalized excess variance is often used as a simple measure of the
intrinsic variability amplitude in light curves,
\citep[see e.g.][]{Nandra:etal:97}:
\begin{equation}
\sigma^2_{\rm NXS} = \frac{1}{N\langle f\rangle^2}\Sigma_{i=1}^{N}[(f_i-f_i^{\rm aver})^2 - \sigma_i^2].
\end{equation}
Here $N$ is the number of data points in the
corresponding time interval,
$f_i$ is the flux of the individual data points, $\sigma_i$ -- their
uncertainty, $f_i^{\rm aver}$
is the smoothed evolution of the flux obtained by a polynomial fit to
the light curve (upper panel of Fig.\,\ref{fig:nv}), and
$\langle f\rangle$ is the mean value of the flux within the interval.
The normalized variance $\sigma^2_{\rm NXS}$
in our case represents the amplitude of
intrinsic flux variations superimposed on the smoothed flux
development. The term $\sigma_i$ under the summation ($\Sigma$) eliminates
the contribution of the Poisson noise.
The bottom panel of Fig.\,\ref{fig:nv} represents the normalized
variance as a function of time during the outburst of EXO\,2030+375.
It can be seen that the amplitude of the variability is high
at the rising phase (corresponding to the flaring episode). Then it decreases
towards the maximum of the outburst and remains low during the decay.
To study periodicity of the flares, we performed a formal period search
in the rising part of the outburst (between MJD 55518.5
and 55521.5) using the Lomb-Scargle periodogram.
The results are presented in Fig.\,\ref{fig:period}. The periodogram shown
in the top panel indicates a clear peak around $\sim$0.3\,days ($\sim$7\,hours).
The averaged profile obtained by folding the light curve with this period
is shown in the bottom panel. The profile shape is
asymmetric and characterized by a steep rise and a slower decay.
\begin{figure}
\resizebox{0.95\hsize}{!}{\includegraphics{period.ps}}
\caption{\emph{Top}: The Lomb-Scargle periodogram of the ``flaring'' part of
EXO\,2030+375 light curve (between MJD 55518.5 and 55521.5).
The peak around $\sim$0.3\,d is clearly seen.
\emph{Bottom:} The ``flaring'' part of the light curve folded with best period
found from the periodogram (0.293\,d).
}
\label{fig:period}
\end{figure}
We used the INTEGRAL data to study the pulse period behavior and pulse
profiles during the outburst. The photon arrival times were converted
to the reference frame
of the solar system barycenter and corrected for the binary orbital
motion using the ephemeris by \citet{Wilson:etal:08}. Using the
pulse-phase-connection technique
\citep[e.g.][]{Staubert:etal:09},
we found the pulse period $P=41.31516(2)$\,s at the
epoch $T_0$(MJD)$=55526.056994$ and the period derivative $\dot P
=-1.9(1)\times 10^{-9}$\,s/s, that indicates significant spin-up. We used
the measured pulse ephemeris to construct and study the pulse
profiles of the source. We could not find any
difference between the profiles obtained during the rise and decay of
the outburst. Figure\,\ref{fig:pp} shows the profiles accumulated
during the entire outburst.
\begin{figure}
\resizebox{0.9\hsize}{!}{\includegraphics{pp.eps}}
\caption{Pulse profiles of EXO\,2030+375 obtained with ISGRI (top) and
JEM-X (bottom) instruments during the outburst (MJD 55518.5 -- 55533.0).
}
\label{fig:pp}
\end{figure}
\section{Spectral analysis\label{sec:spe}}
For the spectral analysis we used JEM-X data between 3.5 and 35\,keV,
and ISGRI data between 20 and 80\,keV. We added systematic uncertainties
at the level of 2\% to the \textit{JEM-X} spectra and 1\% to the
\textit{ISGRI} spectra based on the recommendations of the
instrument teams and the Crab observations.
The spectrum of the source during the outburst
(accumulated between MJD 55518.5 and 55533.0) was modeled with the
cutoff-powerlaw model ($F(E)\propto E^{-\Gamma}\times\exp{[E/E_{\rm fold}]}$,
where $E$ is the photon energy, $\Gamma$ and $E_{\rm fold}$ are the photon
index and the folding energy, respectively) modified at lower energies
by photoelectric absorption.
The best-fit parameters are
$\Gamma=1.6(1)$, $E_{\rm fold}=30(2)$\, keV, the absorption column density
$n_{\rm H}=11(1)\times 10^{22}$
hydrogen atoms per cm$^2$. The uncertainties in
parentheses refer to the last digit and are at the 90\% confidence level.
The value of $n_{\rm H}$ is substantially higher than
measured in previous observations of the source, including
older INTEGRAL measurements
\citep[1--3$\times 10^{22}$\,cm$^{-2}$, e.g.][]{Klochkov:etal:07,
Wilson:etal:08}.
We note, however, that studying $n_{\rm H}$ with JEM-X is generally
problematic as the data below $\sim$3\,keV are not available.
The significance of the measured increase in absorption is
therefore questionable.
The
data quality does not permit spectroscopy
of individual flares. To characterize the spectral
behavior of EXO\,2030+375 during the flaring episode and compare
it with the rest of the outburst, we explored the luminosity-dependence
of the source spectrum during the rising (flaring) part and the decay of
the outburst. We grouped the individual INTEGRAL pointings
according to the measured flux
in the 20--80\,keV range. For each group we extracted and analyzed
the X-ray spectrum using the spectral model described above.
The statistics did not allow us to explore the dependence of each
individual spectral parameter on flux. We therefore fixed $n_{\rm H}$ and
$E_{\rm fold}$ to their averaged values ($11\times 10^{22}$\,cm$^{-2}$
and 30\,keV, respectively) and explored the photon index $\Gamma$
as a function of flux.
\begin{figure}
\resizebox{0.95\hsize}{!}{\includegraphics{frs.eps}}
\caption{Photon index $\Gamma$ as a function of flux during the
rising (flaring) part of the outburst \emph{(left)} and the decay
phase \emph{(right)}.
One Crab corresponds to $\sim$260 cts/s in the specified energy range.}
\label{fig:frs}
\end{figure}
The resulting dependence is shown in Fig.\,\ref{fig:frs} for the
rise (left) and decay (right) of the outburst. Higher values of $\Gamma$
correspond to a softer spectrum (see the model description above).
The horizontal error bars represent the width of the flux bins.
The vertical error bars indicate $1\sigma$-uncertainties.
While the spectrum apparently tends to get harder with
increasing flux during the rising part of the outburst,
it remains roughly constant during the decay.
To quantify this behavior, we performed linear fits to the data points
and calculated the slopes with the corresponding uncertainties.
The slope is
$(-9.0\pm3.7)\times 10^{-3}$\,(cts/s)$^{-1}$ in the rising part and
$(2.8\pm 4.2)\times 10^{-3}$\,(cts/s)$^{-1}$ in the decay of the outburst
(uncertainties at 1$\sigma$ confidence level).
\section{Discussion}
A direct comparison of the flares presented here with those observed
with EXOSAT in 1985 revealed significant similarity.
The peak fluxes and the relative amplitude of the flares are roughly
the same in both episodes. Also the average shape -- fast rise / slow
decay -- is similar in the two cases. The mean period is, however,
different: $\sim$4\,hr for the EXOSAT flares and $\sim$7\,hr for the
INTEGRAL ones. The location of the
EXOSAT flares with respect to the nearest ``normal'' outburst
is difficult to reconstruct because no
monitoring of the source flux (apart from the EXOSAT data themselves) was
performed at the time.
Since the orbital phase of ``normal''
outbursts varies significantly with time \citep{Wilson:etal:02},
the extrapolation of the orbital phase ephemeris back to the EXOSAT
observations would not resolve the problem.
X-ray flares in accreting pulsars are usually attributed to one of the
following mechanisms: (1) instabilities of the accretion flow around/within
the magnetospheric boundary \citep[e.g.][]{Moon:etal:03,Postnov:etal:08},
(2) highly inhomogeneous stellar wind of the donor star
\citep[e.g.][]{Taam:etal:88,
Walter:etal:07}, and (3) nuclear burning at
the neutron star surface \citep[e.g.][]{Levine:etal:00,Brown:Bildsten:98}.
The nuclear burning scenario is, however, difficult to reconcile with
a relatively high accretion rate before and after a flare, which would suppress
the thermonuclear instability \citep{Bildsten:Brown:97}.
In the case of EXO\,2030+375, inhomogeneities of the companion's stellar
wind are also unlikely to be a direct cause of the flares for the
following reasons. First, the viscous time of the accretion disk
that is believed to be present in EXO\,2030+375 during normal outbursts
\citep{Wilson:etal:02} and to even survive during quiescence
\citep{Hayasaki:Okazaki:06}, is at least several days, which would
smooth out any variations in the mass accretion rate $\dot M$ caused by
inhomogeneity of the wind shorter than this time. Second, nonuniform
stellar wind cannot explain the observed quasi-periodic appearance
of the flares. On the other hand, various kinds of
magneto-hydrodynamic instabilities
at the inner edge of the accretion disk may easily lead to oscillations
in the mass flow towards the polar caps of the neutron star
\citep{Apparao:91,Postnov:etal:08,DAngelo:Spruit:10}, leading to the
observed flaring activity. For example, \citet{DAngelo:Spruit:10} have
illustrated that when the magnetospheric radius $r_m$ (where magnetic
field of the neutron star truncates the accretion disk) is larger but close
to the corotation radius $r_c$ (where the Keplerian frequency is
equal to the accretor's spin frequency), matter in the inner regions
of the disk will pile up leading to an increase in the local gas
pressure and, therefore, a decrease in $r_m$. When $r_m$ crosses
$r_c$, the accumulated reservoir of gas is accreted by the neutron
star and the cycle repeats.
To assess the applicability of this scenario to the flares observed in
EXO\,2030+375, we estimated the expected characteristic time scale
of $\dot M$-oscillations. Without going into physical details of the
disk-field coupling at $r_m$ one would generally expect that the
oscillations in $\dot M$ take place on the time scale close to
the local viscous time at $r_m\sim r_c$. Following the standard
$\alpha$-viscosity prescription \citep{Shakura:Sunyaev:73}, this time
can be estimated as $\tau_c\sim r_c^2/\nu(r_c)\sim
1/[\Omega\alpha(H/R)^2]$, where $\nu(r_c)$
is the viscosity at $r_c$, $\Omega$ is the spin frequency of the
neutron star, and $(H/R)$ the semithickness of the accretion
disk. Using ``standard'' values of $\alpha=0.1$, $H/R=0.05$, and the
known pulse period $P\simeq 40$\,s, one gets $\tau_c\sim 7$\,hr,
i.e. of several hours, as was observed.
The averaged profile of the flares is characterized by a steep rise and
a slower decay (Fig.\,\ref{fig:period}), which is very similar
to the flares observed with EXOSAT in 1985 (Fig.\,2 of
\citealt{Parmar:etal:89a}). According to the authors, such a shape
suggests a ``draining reservoir'' that is in line with the picture
described above (matter piling up on the inner edge of the disk).
The observed difference in the spectral behavior between the
flaring part and the rest of the outburst (Sect.\,\ref{sec:spe})
suggests different configurations of the region where matter
couples to the field lines. Such a difference is indeed expected
if the flares are caused by the oscillating inner edge of the accretion
disk. In this case, matter from the oscillating inner disk rim
would couple to different dipole field lines
of the neutron star (and follow them) compared to the decay part of
the outburst where the configuration of the inner disk rim
is presumably stable.
The difference in the mean period of flares in the INTEGRAL
and EXOSAT observations can also be understood in the described picture.
The period must depend on the mass transfer rate through the accretion
disk, i.e. time needed to refill the reservoir. This rate could be
different between the EXOSAT and INTEGRAL observations due to,
e. g., changes in the state of the Be-disk.
In Fig.\,\ref{fig:lcflares}
one might also notice some shortening of the the flare separation
time as the flux increases (although this behavior is difficult to
quantify with the available statistics). Such behavior, if real,
might reflect shortening of the reservoir refill time as the mass
transfer rate increases towards the maximum of the outburst.
Thus, we argue that the observational appearance of the flares
in EXO\,2030+375 suggests that the instability of the inner disk edge
(pile-up/draining of matter) is the most probable cause of the flares.
It is important to note that
the rarity of the detected flaring episodes
(even considering the relatively sparse observational coverage)
means that the range of physical conditions needed to initiate
flares could be very narrow, which would
lead to the serendipitous character of the phenomenon.
\begin{acknowledgements}
The work was supported by the Carl-Zeiss-Stiftung and by DLR grant BA5027.
This research is based on observations with INTEGRAL, an ESA
project with instruments and science data centre funded by ESA member
states.
The authors thank the anonymous referee for useful suggestions.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Urban Advantages and Disadvantages}
More than half of the world population lives in cities~\cite{Cohen:2003}. With 180,000 people moving to cities every day~\cite{Intuit:2010}, urban population is expected to grow, reaching $70\%$ of the global population by 2050~\cite{Cities:2010,Roberts:2011}.
There are several advantages of urban settlements, such as less energetic requirements per capita, higher incomes, innovation, and productivity~\cite{Glaeser:2011,Bettencourt:2007,Bettencourt:2010}. In spite of---or perhaps because of---being highly attractive for people, modern cities also face several problems, such as mobility, crime, disease, pollution, and other social problems.
There have been several proposals concerning every urban problem, with different degrees of success. There are cities where the major problem might be mobility (Mexico City, Beijing~\cite{Gyimesi:2011}), safety (Ciudad Ju\'arez, Baghdad), unemployment (Detroit, Madrid), segregation (Chicago, Pretoria), traffic accidents (El Cairo, Dar-es-Salaam), or lack of infrastructure (Lagos, Kabul). Since there are different causes for different problems, there will be no single solution for all urban problems: several solutions have to be explored in parallel.
Urban planning has been guiding the development of cities for decades, at least in developed countries. Planning is certainly useful: it is better to deal with situations before they become problematic. However, urban planning has been rigid so far: how can future requirements be predicted as cities grow and embrace new technologies and customs? Just like a century ago cities were not planned to use cars as major means of transportation, cities cannot be planned now for their requirements of the next fifty years. Moreover, it is only recently that researchers have been able to develop descriptive models of urban growth~\cite{Andersson:2002b,Andersson:2002,Yamins:2003,Silva:2005,Mahiny:2012}.
The limitations of urban prediction are due to the complexity of cities~\cite{GershensonHeylighen2005,HeylighenEtAl2007,Gershenson:2011e}. Complexity implies that the components of a system are not separable. This lack of separability is due to relevant \emph{interactions} between components: The future state of components is co-determined by interactions, which cannot be enumerated, ordered, or predicted. Thus, prediction from initial and boundary conditions is limited.
Cities can usefully be described as complex systems~\cite{portugali2000self,batty2005cities,Batty:2008}, since their components interact and co-determine their future. Thus, urban planning is limited~\cite{White2000High-resolution} by the very nature of their complexity. This does not imply that general properties of cities cannot be estimated, but that precise prediction is hopeless. To complement this lack of prediction, living technology can serve cities by providing a greater degree of adaptability and robustness.
In the next section, an overview of living technology and its properties is given. Section \ref{sec:urbanProblems} provides an extensive (although non-exhaustive) description of urban problems and solutions offered by living technology. In particular, problems in mobility, logistics, telecommunications, governance, safety, sustainability, and society and culture are discussed. Section \ref{sec:how} offers guidelines to develop urban living technology, using public transportation systems as a case study.
The paper concludes with a discussion on the usefulness of describing cities as living systems.
\section{Living Technology}
The term ``living technology" has been used to describe technology that is based on the core features of living systems~\cite{Bedau:2009}. Living technology is adaptive, learning, evolving, robust, autonomous, self-repairing, and self-reproducing.
\emph{Adaptation}~\cite{Holland1975,Holland1995} can be described as a useful change in a system in response to changes in its environment~\cite[p.19]{GershensonDCSOS}. Living systems are constantly adapting because their environment is dynamic. Adaptive technology is necessary where problems are dynamic. Certainly, there are different degrees of adaptation: a thermostat adapts only to changes of temperature, while an autonomous car has to adapt to changes in roads, traffic states, other vehicles, behavior of drivers, etc.
\emph{Learning} and \emph{evolution} can be seen as a second order adaptation, since they imply a permanent change in a system. In other words, after learning or evolution, a system will respond in a different way to similar circumstances. Learning and evolution occur at different timescales: learning is a type of adaptation within a lifetime, while evolution is a type of adaptation across generations. Learning and evolving technologies are useful because they can adapt to novel circumstances. With these properties, the same system will be able to function in a broader range of situations. This increases the potential variety and complexity that the system can cope with~\cite{Ashby1956,BarYam2004,Gershenson:2010a}.
A system can be said to be \emph{robust} if it continues to function in the face of perturbations~\cite{Wagner2005}. Robustness---also called resilience---is prevalent in living systems and desired in technology~\cite{vonNeumann1956,Jen2005}, as it complements adaptation by allowing a system to ``survive" changes in the environment before it can adapt to them. Robustness and adaptation are deeply interrelated, since they are different ways to cope with unpredictable environments. Robustness is passive (changes are resisted by the system), while adaptation is active (changes cause a reaction in the system). Robustness can be promoted by different properties~\cite{Gershenson:2010}, such as modularity~\cite{Simon1996,Watson2002,Schlosser:2004,Callebaut:2005,BalpoGershenson:2011}, degeneracy
\cite{Edelman:2001,FernandezSole2003,Wagner2004,Whitacre:2010}, and redundancy~\cite{GershensonEtAl2006}.
\emph{Autonomy}~\cite{Barandarian2004,MorenoRuiz2006,KrakauerZanotto2007} implies a certain independence of a system from its environment. Adaptation and robustness are requirements for autonomy, since they enable a system to withstand perturbations. Additionally, the autonomy of a system implies a certain degree of control over its own production~\cite{VarelaEtAl1974,McMullin2004} and behavior~\cite{Gershenson:2007}. Living systems have a high degree of autonomy. Technology has a tendency to become more and more autonomous of humans: from robots~\cite{Bekey:2005} to trading algorithms in stock markets~\cite{DeMarzo:2006}. This enables technology to respond to changes at faster rates. However, autonomous technology is also generating faster changes that affect other technologies.
\emph{Self-repair} and \emph{self-reproduction} can be seen as particular cases of \emph{self-organization}~\cite{GershensonHeylighen2003a}. Almost any system can be said to be self-organizing~\cite{Ashby1962}. However, it is \emph{useful} to describe a system as self-organizing when one is interested in relating how the interactions of elements affect the global properties of a system. This can be applied to living systems at several scales. For technology, self-organization can be used as an approach to build adaptive and robust systems~\cite{GershensonDCSOS}: interactions are designed so that elements find solutions by themselves. Thus, systems can adapt constantly to changes in their environment.
There cannot be a sharp distinction between non-living and living technology (just as there cannot be a sharp distinction between non-living and living systems). Nevertheless, it can be said that technology will be ``more living" as it has more and more of the core properties of living systems.
Living technology can be distinguished as primary or secondary~\cite[p. 91]{Bedau:2009}. \emph{Primary} living technology is constructed from non-living components, while \emph{secondary} living technology depends on living properties already present in its elements. Cities are secondary living technology, since living systems (humans, animals, plants, bacteria) are part of urban spaces. Nevertheless, the non-living components of cities have been acquiring with technology certain aspects of living systems, as mixed networks of soft, hard, and wet ALife~\cite[p. 92]{Bedau:2009}.
If cities always included living systems, have they always been using living technology? The answer depends on the deep question of the definition of life, which is far from being settled. To be able to decide ``how living" a system is, measures based on information theory can be used~\cite{Gershenson:2007}: we can measure how much the information of a system depends on the information of its environment. In this sense, ``more living" systems are those which are more autonomous from their environment, i.e. they produce more information about themselves than the information about themselves produced by their environment. Still, this measure depends on the scale at which the information is measured~\cite{GershensonFernandez:2012}. For example, it can be argued that a bacterium is more autonomous than a cell from a multicellular organism because it produces more of its own information. However, a multicellular organism produces more information about itself than a bacterial colony, since its organization at the multicellular scale can maintain its own integrity to a larger degree than the bacterial colony~\cite{Gershenson:2010a}. Thus, it can be argued that the organism is more autonomous than the colony at the multicellular scale.
If we are interested on deciding ``how living" urban technology is, we have to measure how much an urban system is able to produce its own information, which reflects its organization and thus control over its own dynamics, \emph{at the urban scale}. Different urban systems can be composed by the same living and non living components, e.g. traffic (drivers, pedestrians, vehicles, traffic lights, etc.). But different organizations of the urban system, e.g. traffic light coordination methods, will deliver different informational measures for the system, which will reflect their abilities to adapt, learn, evolve, and self-repair. For each urban system, if we increase its ``liveness" with living technology, the system will be able to deliver a better performance in comparison with a system without the properties of living systems.
\section{Solutions for Urban Problems}
\label{sec:urbanProblems}
Cities have been described metaphorically as organisms, e.g.~\cite{Dawson:1926,Spilhaus:1969}: they grow, have a metabolism, an internal organization, transportation networks of matter, energy, and information, and telecommunications have been characterized as ``nervous systems". Urban areas also reproduce and repair themselves, although their mechanisms are more akin to grasses than animals. Even thermodynamically, cities take matter, energy and information from their environment, transform them, and produce waste to maintain their organization, just like living systems.
However, Lynch~\cite{Lynch:1981} argued that descriptions of modern cities as living organisms or as machines are inadequate, even when they contain all twenty subsystems required by living systems, as defined by Miller~\cite{Miller1978Living-Systems}.
Still, the promise of living technology towards improving urban systems and thus transforming the nature of cities was not yet considered three decades ago. Moreover, Batty~\cite{Batty:2012Cities} has recently argued that the scientific study of cities is transitioning ``from thinking of `cities as machines' to `cities as organisms'".
Bettencourt et al.~\cite{Bettencourt:2007} discovered that---in spite of several similarities---various properties of cities belong to different universality classes than those of biological organisms. Nevertheless, similar to living organisms, cities are constantly adapting~\cite{Bettencourt:2010}. In any case, this paper is not focussed on deciding whether cities are usefully described as living systems or not, but on exploring the use of living technology to solve urban problems.
Traditional approaches are efficient for \emph{stationary} problems, i.e. a solution is found, implemented, and the problem is solved. However, most
urban problems are \emph{non-stationary}~\cite{Forrester1969Urban-Dynamics,Batty1971Modelling-Citie,Gershenson:2011b}: population changes with years, opinions can change within days, energy, resource, and waste requirements change with the seasons and with the hours of the day, traffic changes every second. Not only there are changes occurring constantly in urban spaces, but these occur at different scales. Solutions to these problems have to be robust and adapt, \emph{matching the scales} at which changes take place~\cite{GershensonDCSOS,Gershenson:2011a}.
Since urban problems are dynamic, urban technology has to find new solutions as problem changes by adapting, learning, and evolving. Living technology can offer this type of solutions~\cite{Mehaffy:2011,Alexander:2003}. Moreover, cities have been invaded by information technology~\cite{Kitchin:2011}, becoming a mesh of sensors, actuators, and controllers, exploiting the combined abilities of citizens and technology.
Biourbanism~\cite{williams1997biourbanism} has already proposed a similar path, looking at interdependencies between all the components of urban systems, focussing on sustainability and ecology. Biourbanism proposes the use of technologies that are closer to biology with the aim of having a reduced impact on the environment.
Information technology (IT) is bringing several properties of living systems to urban spaces~\cite{Kitchin:2011}. IBM's smart cities program aims at solving some urban problems with the aid of IT~\cite{Dodgson:2011,Harrison:2011}. The FuturICT european flagship project~\cite{Helbing:2011} proposes the integration of techniques from several disciplines to solve global problems, many of them urban. The Earth 2.0 project\footnote{\url{http://earth2hub.com}} is also proposed at a global scale, using IT to build more adaptive and sustainable global and urban systems. The organic computing paradigm~\cite{OrganicComp:2011} focusses on information processing systems with properties of living systems. Organic systems can be considered as living technology.
In the next subsections, several urban problems and potential solutions with living technology are presented.
\subsection{Mobility}
The movement of people and goods is one of the major urban problems. It requires expensive infrastructure (roads, rails, ports, stations, bridges, vehicles, fuel, signalization). When mobility is inefficient or saturated, people lose time and money, gain stress, and more pollution is generated. Overall, the quality of life is reduced when mobility is limited or not efficient. There are several problems within urban mobility, so there will be no single solution for all of them~\cite{Cairns:2004}. At least eight interrelated aspects of urban mobility can be identified:
\begin{description}
\item[Transportation requirements. ] There is no mobility problem if people and goods do not have to be displaced. It is not possible for everyone to study, work and grow produce at home, but many actions can be taken to reduce the need of moving people and merchandise, i.e. the mobility demand.
\item[Scheduling. ] Congestion occurs when there are too many people in the same place at the same time. If people can transport themselves with more flexible schedules, then the demand of rush hours can dissipate over longer periods of time.
\item[Quantity. ] Too many vehicles or people saturate roads and public transportation systems. To reduce this, some cities use measures to demotivate use of private vehicles, such as high taxes, congestion charges, and limited parking. More flexible approaches to reduce vehicle quantity are carpooling and carsharing~\cite{Gansky:2010}, e.g. Zipcar and Buzzcar.
\item[Capacity. ] Building more and broader freeways, bike lanes, public transportation systems and efficient traffic lights increases the capacity of urban mobility. An increased capacity can be expensive, although technology can allow for increases in capacity at reduced costs.
\item[Behavior. ] Inadequate behavior of drivers or passengers can lead to delays in transportation. Examples for drivers include speeding, compulsive lane changing, and texting while driving. Examples for passengers include pushing and blocking, which can occur in different circumstances.
Potential interventions for restricting inadequate behaviors and promoting positive behaviors include education campaigns, fines, real-time information and social participation~\cite{singhal2008performance}.
\item[Infrastructure and technology. ] Infrastructure such as freeways, public transportation, bike lanes, and vehicle sharing systems can contribute to improve mobility. Technology can complement infrastructure, by enhancing its capacity. For example, traffic sensors can be used to coordinate traffic lights, avoid traffic jams, and suggest alternative routes.
\item[Society. ] In most societies, owning a car implies prestige, reflecting certain economic success. However, people are becoming so successful that roads are saturated. In several cities, people naturally prefer alternative modes of transport. With a social acceptance of car-owning alternatives, it will become easier to balance different modes of transportation farther from private cars.
\item[Planning and regulation. ] Even when urban planning has limitations, cities suffer when there is no urban planning at all. In many cities this is complicated because politicians and not urbanists make the decisions on urban projects. Also, some cities do have planning and projects, but there is no enforcement nor regulation. Thus, plans never materialize and projects never are implemented.
\end{description}
There are different actions that can be taken to improve different aspects of the eight factors mentioned above. For example, more capacity can be built. But if the quantity increases faster than the capacity, the improvement will be severely limited and problems will not be solved. In general, all of the eight factors have to be considered in parallel to improve urban mobility. In the next sections, examples of potential applications of living technology to address different problems in urban mobility are presented.
\subsubsection{Public Transportation}
When thousands or even millions of people have to move in urban areas through similar routes, mass transit becomes a better alternative than private motorized vehicles. Metro, bus rapid transit (BRT), trams, buses, and trains have been used since the nineteenth century for this purpose.
According to theory, passengers arriving randomly at stations wait the least when headways---the temporal interval between vehicles---are equal~\cite{Welding:1957}. However, this configuration is always unstable, for all public transportation systems~\cite{GershensonPineda2009}. Random arrivals at stations will cause some stations to be busier than others. When a vehicle arrives at a busy station, it might be slightly delayed, increasing the headway with the vehicle ahead and reducing the headway with the vehicle behind. The longer headway might cause further delays at the next station, increasing even more the headway with the vehicle ahead and decreasing even more the headway with the vehicle behind. This equal headway instability leads to the formation of ``platoons" of vehicles that affect negatively the service, leading to long delays for passengers. There have been several approaches for dealing with equal headway instability in particular transportation systems~\cite{Turnquist:1980}.
Recently, it was found that transportation theory had misguided assumptions for decades~\cite{Gershenson:2011a}, namely that vehicles along a route will have the same travel time, thus developing methods that aim at maintaining equal headways, reducing waiting times for passengers \emph{at stations}. However, in order to maintain equal headways, some vehicles have to idle at stations. A self-organizing method was proposed~\cite{Gershenson:2011a}, where the equal headways are relaxed and even when passengers wait more at stations, the total travel times are reduced by a slower-is-faster effect~\cite{Helbing:2000,Helbing:2009}. The proposed method uses ``antipheromones" to make local decisions depending on neighboring vehicles and passenger demands at current stations, adapting to changing demands and delivering a ``supraoptimal" performance. The details of this solution are discussed as a case study in Section~\ref{sec:antiph}.
\subsubsection{Traffic Lights}
The coordination of traffic lights is an exponential-complete problem~\cite{PapadimitriouTsitsiklis1999,Lammer:2008}. Moreover, the traffic configuration changes constantly, as demands at intersections vary at the seconds scale. For this reason, fixed, optimizing approaches are limited for traffic light control~\cite{GershensonRosenblueth:2010}.
Adaptive methods, some of which are biologically-inspired, have been proposed to regulate traffic lights. Faieta and Huberman~\cite{FaietaHuberman1993} proposed an algorithm inspired in firefly synchronization, while Ohira~\cite{Ohira1997} proposed a controller based on an analogy with neural networks.
Self-organizing traffic lights~\cite{Gershenson2005,HelbingEtAl2005,CoolsEtAl2007,Lammer:2008,Prothmann:2009,GershensonRosenblueth:2011,deGier:2011} can adapt to the local traffic demand, leading to an emergent and robust global coordination of traffic lights. Some of these methods are in the process of being implemented~\cite{Lammer:2010}, reporting considerable improvements in waiting times for cars, pedestrians, and public transport. This leads to economic, energetic, environmental, and social savings.
\subsubsection{Real-time Information}
\label{sec:RTinfo}
The commercialization of GPS devices allowed drivers to query for the shortest route to their destination. However, once several people were using GPS, shortest routes were saturated, since everyone was advised to follow them. Shorter but not fastest. Real-time information---available for decades in radio traffic reports---can help drivers adapt their route according to the current traffic situation. A limitation of radio reports is that they are broadcasted: all drivers get the same information, most of which might not be relevant, and drivers cannot demand for particular information.
This scenario has changed in recent years, with applications such as Google Maps\footnote{\url{http://maps.google.com}} and Waze\footnote{\url{http://www.waze.com}}, which provide real time traffic information on demand.
A key element of real-time information systems consists of sensors~\cite{Chong:2003,dressler2007selforg}. Once traffic states are detected, broadcasting or making them available is relatively straightforward. Since there are different types of sensors (fixed, mobile),
sensor integration~\cite{Qi:2001} is a relevant problem to obtain useful information.
Intervehicle communication can provide useful real-time local information, which can be exploited to adapt to dynamic traffic states and improve traffic flow~\cite{Kesting:2008}.
Real-time information for public transportation systems can also help passengers to adapt their routes more efficiently, and even their behavior~\cite{GershensonPineda2009}.
In general, location-based services offer a broad application potential~\cite{Ratti2006Mobile-Landscap}.
\subsection{Logistics}
\label{sec:logistics}
Biologistics~\cite{Helbing:2009a} notes that the organization, coordination and optimization of various material flows is not restricted to artificial systems, but that living systems also have to deal with material flows. Moreover, living systems can handle material flows efficiently, adaptively, robustly, and learn from past experiences. Thus,
with biological inspiration, using principles of modularity, self-assembly, self-organization, and decentralized coordination, artificial logistic systems can be designed that can adapt efficiently to changes of demand.
A drawback to traditional approaches in logistics is that the supplies and demands for different goods are dynamic and unpredictable. This demands approaches where systems can adapt to changing demands at the same scale at which changes occur.
For example, swarm intelligence~\cite{BonabeauEtAl1999,Kennedy:2001,Trianni-Tuci:09:ecal} has been applied to several problems in logistics~\cite{Svenson:2004}.
Computationally, algorithms inspired by swarms or neurons are equivalent~\cite{Gershenson:2010b}, since they function at multiple scales, allowing them to compute solutions to problems at a faster scale and at the same time adapt to changes in problems at a slower scale. This is a desired property in logistics and many other areas.
\subsection{Telecommunications}
A distinction can be made between synchronous and asynchronous communicaiton~\cite{DesanctisMonge1999,GershensonSOBs}. IT has reduced delays of information transmission, allowing for technologies with faster response rates. Moreover,
IT has made it possible to shift from from broadcasted information to information on demand. Availability of information is a requirement for living urban technology, since relevant information is required in order to adapt, learn, and evolve. This was already illustrated in Section \ref{sec:RTinfo}.
Telecommunications have an essential role towards the use of living technologies in urban spaces. Not only for information transmission among citizens, but also among devices and systems~\cite{Resch:2011}. For this purpose, several approaches have been proposed to build adaptive, flexible and robust telecommunication networks~\cite{dressler2008bio-inspired,dressler2010bioinspired}. These networks are becoming so complex and operate at such speeds, that their technology can only function efficiently exhibiting the properties of living systems.
Telecommunication systems are relevant not only for transmission of information, but enable other uses of living technology in urban spaces, such as governance~\cite{pitroda1993development}.
\subsection{Governance}
Bureaucracies are often seen as rigid, slow, and inefficient.
Living technology can enable the adaptive transmission of relevant information to govern cities~\cite{GershensonSOBs}. Any adaptive system requires sensors to be able to detect when changes are required. An obstacle in governance is that sensors are rather poor to allow governments to make informed decisions. Simply there is no infrastructure to detect what are the requirements of citizens. For example, India is connecting 250,000 local governments (Panchayats) to deliver and obtain information to and from citizens~\cite{Panchayats:2010}.
Sensors are important but not the only aspect where changes are being made. Technology can also be used to make better collective decisions~\cite{RodriguezSteinbock2004,RodriguezEtAl2007}. This possibility enables societies to respond adaptively to different situations. This also helps governments to better administer cities.
Governments have also been making their data publicly available, so that citizens can use this information in novel ways~\cite{Bizer:2009}. Opening data and information enables many potential applications.
Also data created by citizens can be useful. For example, after the 2010 Haiti earthquake, people used OpenStreetMap\footnote{\url{http://www.openstreetmap.org/}} to improve maps and assist rescue and humanitarian aid efforts, identifying via satellite pictures collapsed buildings, refugee camps, and other damages.
The availability and processing of masses of urban data open the potential to governments that adapt constantly to changes in demand of their citizens. Moreover, they allow an increased citizen participation in governance, slowly fading the differences between governors and the governed. An extreme democracy can be reached when the opinion of every citizen has the same weight on any political affair. This would be achieved only with living technology, since such a system would have to adapt constantly to the changes in the population.
\subsection{Safety}
In a similar way that living technology can improve governments, it can improve urban safety. On the one hand, prompt and adaptive response to natural and artificial catastrophes is facilitated. On the other hand, an urban mesh of sensors can increase public safety by monitoring public and private spaces, thus increasing citizen accountability. Simply having cameras to detect traffic infractions enforces people to comply with traffic rules, which---if designed properly---lead to increased road safety.
Artificial immune systems (AIS) have been proposed to to prevent intrusions in networked systems~\cite{hofmeyr1999immunity}. AIS exhibit properties of their biological counterparts: they are distributed, robust, dynamic, diverse and adaptive. Since intrusions are seldom repeated, security systems have to be flexible enough to adapt and respond to novel situations constantly.
If used properly, living technology could also reduce crime rates. Having an effective police is not a solution for urban crime, since its causes seem to lie in unemployment, lack of opportunities, social influence, and several other factors~\cite{weatherburn2001causes}. Nevertheless, crime prevention is necessary, and it will be more effective if it exhibits properties of living systems~\cite{ekblom1999can}, since changing circumstances, trends and behaviors open constantly new niches for crime. Thus, an effective crime prevention has to adapt to these changes, to learn from previous experiences, and to be robust in the process. It might be just a coincidence, but life has become safer as technology has evolved~\cite{pinker2011better}. The causal relations between technology and safety have yet to be explored, but this trend probably will continue, increasing safety as technology becomes ``more living".
\subsection{Sustainability}
Sustainability is the capacity to endure. For cities, sustainability involves not only environmental relations, but also economical and social. Material and energetic resources are required to ``fuel" cities, as well as economic and social benefits to attract and sustain citizens \cite{Trantopoulos2012Toward-Sustaina}.
Concerning material sustainability, pollution has to be considered. If there is less waste produced, then the complexity of waste management will be reduced. Cleaner and more efficient technologies can help in this direction. For example, if traffic flow is more efficient, less pollution will be produced by motor vehicles. Also,
local production reduces transportation and transmission burdens, but the cost of production might be higher. Thus, a balance between mass production (cheaper to produce but distribution required) and local production (more costly to produce, cheaper to distribute) should be sought. Nevertheless, living technology can contribute to both reducing the cost of local production and to increase the efficiency of distribution (See Section \ref{sec:logistics}).
Synthetic biology~\cite{Benner:2005} (wet second-order living technology) is promising for producing cleaner fuels~\cite{Lee:2008}, as well as technology to reduce or prevent pollution, such as buildings that absorb carbon dioxide and bioluminescent trees that do not require electricity~\cite{Armstrong:2010}.
The efficient and adaptive production and distribution of energy, as envisioned by the concept of a ``smart grid"~\cite{gellings2009smart,Anghel:2007} is similar to other urban problems: there is a varying demand, as well as a varying production, which ideally should match the demand. Living technology can certainly benefit energy grids, coordinating local generation of energy and distributing it ``on demand".
Another application of living technology is the dynamic regulation of rainwater to collect water and prevent floods, where catchment systems react on the weather forecasts and water supply levels~\cite{Mims:2011,Ruhnke:2011}.
``Smart skins" for buildings have also been proposed for temperature regulation minimizing energy consumption.~\cite{Ritter:2007}.
A sustainable economy should produce more than what it consumes. Moreover, it has to accommodate employments, opportunities, and pensions for dynamic populations (aging in some countries, growing fast in others). W. Brian Arthur has recently described ``the second economy"~\cite{Arthur:2011}, based on information technology, where processes are interacting, adapting, and having an effect on the ``physical" economy. Arthur mentions that the second economy has properties of living systems, since digital devices and processes are starting to sense, compute, make decisions, and perform actions adaptively and independently of humans.
Businesses and enterprises also have to develop and acquire living technology, since the demands of the markets are changing constantly and at increasing speeds. Organizations that are adaptive and robust will have better chances of enduring unpredictable changes in the economy. Moreover, urban living technology is itself a novel business niche~\cite{Arup:2011}.
Living technology can also have a positive effect in the social aspects of urban spaces. Safety was already mentioned, but in general living technology can help citizens to be more cooperative. Taking the example of driving, in some cases it might be beneficial for a driver to drive in such a way that affects negatively other drivers, tempting them to do the same. When a few drivers follow this behavior, the traffic becomes worse for everybody, including those that attempt to get a benefit.
Cooperation has been extensively studied with game theory~\cite{Axelrod:1981,Nowak2006}. Living technology can provide several alternatives to promote cooperation. On the one hand, those who do not cooperate could be punished automatically. On the other hand, those who do cooperate could be rewarded. Moreover, living technology could help change situations in such a way that it will be beneficial for individuals to behave in such a way that is beneficial for the society as well. In other words, if the payoff for cooperating is always the highest, there will be no social dilemmas: everybody will selfishly cooperate.
\subsection{Society and Culture}
One example of a social benefit is given with innovation, which is already promoted by cities~\cite{Bettencourt:2007}. Can living technology accelerate innovation in cities? It seems that the answer is affirmative, at least indirectly: if living technology can solve at least some of the urban problems mentioned above, it will increase the attractiveness of cities to citizens. Moreover, it will increase the ``carrying capacity" of sustainable cities. Since larger cities tend to be more innovative, and living technology would allow cities to grow even more, it can be concluded that such ``living cities" will have an increased innovation rate. And innovation not only in science and technology, but in culture, education, and art as well.
Since IT and the Internet are reducing the burden of transportation, people are exchanging information remotely and globally, overflowing the benefits of urbanization across cities.
Social media---such as Twitter and Facebook---are transforming and facilitating social interactions. For example, ``Social moods" have been detected~\cite{Bollen:2011}. Technology over social networks could potentially be used to steer social behavior, for example preventing unhealthy habits and promoting healthy ones~\cite{Gershenson:2011c}.
\section{How to do it?}
\label{sec:how}
In the previous section, examples of existing and potential urban living technologies were mentioned. This section will focus on how living technology can be applied to urban problems.
Recently, a methodology was developed for designing and controlling systems that require to be adaptive and robust using the concept of self-organization~\cite{GershensonDCSOS}. Instead of designing a system to solve a problem that is changing constantly, with self-organizing systems components are designed so that they find solutions by interacting among themselves. This allows them to \emph{autonomously} \emph{evolve}, \emph{learn} and \emph{adapt} to changes in the problem and to continue functioning in a \emph{robust} way. The methodology focuses on identifying the nature of interactions and eliminate or reduce negative interactions (``friction") and promote positive interactions (``synergy"~\cite{Haken1981}). Interaction improvement always leads to system performance improvement~\cite{GershensonDCSOS}. This approach is useful when the problem or situation is unknown, undefined, or dynamic.
The methodology is only one of several that have been proposed with similar aims in the literature. A review and comparison can be found in~\cite{Frei:2011}. Engineering methodologies that embrace complexity are promising for developing living technology. This is because they offer frameworks where artificial systems with the properties of living systems can be developed.
In the next subsection, public transportation systems are used as a case study where living technology based on self-organization offers even better performance than the theoretical optimum.
\subsection{A case study: self-organizing public transportation systems}
\label{sec:antiph}
Passengers arriving randomly at stations will wait the least time if the headways (intervals between vehicles) are equal~\cite{Welding:1957}, as illustrated by Figure \ref{fig:headways}.
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=4in]{img/headways}
\end{center}
\caption{
A. Equal headways lead to shorter passenger waiting times at stations. On average, the waiting time at stations is half the intervehicle interval. B. With unequal headways, passengers also are expected to wait half the current intervehicle interval, but there is a higher probability of passenger arrival within longer intervals, leading to higher average waiting times~\cite{Gershenson:2011a}.
}
\label{fig:headways}
\end{figure}
Even when an equal headway configuration is desired, this is unstable, as explained in Figure \ref{fig:HeadwayDeviation}. It is like an inverted pendulum, where any perturbation kicks the system off balance and brings the pendulum down. In a similar way, public transportation systems ``prefer" to have unequal headways, as small delays amplify with a positive feedback, leading to the collapse of the system. Much of public transportation engineering for the past fifty years has dealt with trying to force transportation systems into maintaining an equal headway configuration~\cite{Turnquist:1980}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=15cm]{img/HeadwayDeviation}
\caption{Equal headway instability. a) Vehicles with a homogeneous temporal distribution, \emph{i.e.} equal headways. Passengers arriving at random cause some stations to have more demand than others. b) Vehicle $c$ is delayed after serving a busy station. This causes a longer waiting time at the next station, leading to a higher demand ahead of $c$. Also, vehicle $d$ faces less demand, approaching $c$. c) Vehicle $c$ is delayed even more and vehicles $d$ and $e$ aggregate behind it, forming a platoon. There is a separation between $e$ and $f$, making it likely that $f$ will encounter busy stations ahead of it. This configuration causes longer waiting times for passengers at stations, higher demands at each stop, and increased vehicle travel times. The average service frequency at stations is much slower for platoons than for vehicles with an equal headway~\cite{GershensonPineda2009}.}
\label{fig:HeadwayDeviation}
\end{center}
\end{figure}
In this context, a self-organizing method was developed with the aim of not only maintaining equal headways, but also to recover from unequal headway configurations. Following the inverted pendulum analogy, the goal was to build a system that would not only to prevent the pendulum from falling, but also to lift it up starting from a fallen position.
Inspired by the adaptivity of ant communication~\cite{CamazineEtAl2003}, the method was tested and refined. One type of ant communication involves the secretion and sensing of pheromones. For example, if an ant finds a source of food, it will return to its nest with some food while leaving a pheromone trail. Other ants have a tendency to follow pheromone trails, proportional to the pheromone concentration. Thus, if more ants follow the pheromone trail and find the source of food, they will also return to the nest bringing food and reinforcing the pheromone trail, increasing the probability of recruiting more ants. Once the food source is exhausted, ants stop reinforcing the pheromone trail, which evaporates with time to prevent more ants from going to an empty source. Once a new source of food is found by exploring ants, new pheromone trails are formed. This communication via the environment is also known as ``stigmergy"~\cite{TheraulazBonabeau1999}. Functionally, the cognition of insect colonies mediated by stigmergy is analogous to neural cognition~\cite{Gershenson:2010b}.
The self-organizing method proposed for regulating public transportation headways is also stigmergic. However, instead of using pheromones, it uses ``antipheromones". Pheromones are placed by insects and evaporate with time, thus reducing their concentration. Antipheromones are virtual markers that increase their concentration with time, while they are erased by passing vehicles. A simple algorithm determines how much time each vehicle should spend at each station depending on the amount of \emph{passengers} waiting at the station, the antipheromone concentration (which is directly proportional to the \emph{time} since the vehicle ahead departed), and the \emph{distance} to the vehicle behind~\cite{Gershenson:2011a}. This algorithm enables each vehicle to adapt to the demand at each station, preventing idling that occurs when equal headways are maintained, and allows enough robustness to prevent the platooning of vehicles and flexibility to recover from platooned configurations.
Discrete computer simulations were performed to compare the self-organizing method with a ``default" method, which does not restrict any waiting time and always leads to equal headway instability, and an ``adaptive maximum" method~\cite{GershensonPineda2009}, where there is a minimum waiting time at stations for vehicles and a maximum waiting time is modified depending on the global passenger demand and headways are always maintained, but not recovered. In the simulations, vehicles have a maximum passenger capacity and move discretely one space unit per time step, unless there is another vehicle ahead, passengers are boarding or descending at stations, or there is another restriction, such as waiting times at stations. Passengers arrive randomly at stations with a Poisson distribution on average every $\lambda$ time steps. When a vehicle arrives at a station, passengers scheduled to descend exit taking one time step each. Then passengers waiting at the station board taking one time step each until the vehicle is full or leaves the station.
Results for a homogeneous scenario, with equidistant stations and initial positions of vehicles and equal passenger demand ($\lambda$) at stations, are shown in Figure \ref{fig:homo} for four different passenger demands. The headways in the default method collapse (as seen by the high standard deviations of intervehicle frequencies) leading to very high waiting times.
Surprisingly, the self-organizing method, even when headways are not maintained (although the system does not collapse), produced waiting times even lower than those of the maximum method, which maintained equal headways. Theory would tell us that waiting times are optimal for an equal headway configuration, meaning that the self-organizing method delivers supraoptimal performance. Still, when passenger waiting times are separated between total waiting times and waiting times at stations, the maximum method indeed has the minimum waiting times at stations, which is what the theory tells us. However, the theory assumes that travel times are independent of waiting times at stations, and they are not. In order to keep equal headways, some vehicles must idle, while others must leave some passengers behind. The self-organizing method is flexible enough so that headways are not maintained but neither collapsed, while passengers at stations are served on demand. Thus, even when waiting times at stations are higher, the total waiting times are lower.
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=3.5in]{img/allhomo-results2}
\end{center}
\caption{
Results for homogeneous scenario. A. Passenger delays for methods: ``default" (\emph{DF}), ``max" (\emph{MX}), and ``self-organizing" (\emph{SO}), for different passenger demands (lower $\lambda$ means higher demand). Lower boxes at each column show waiting times at stations. Higher boxes show total waiting times. B. Headway standard deviations. Lower $\sigma_f$ implies more regular headways. \emph{DF} shows unstable headways, \emph{MX} equal headways (except for $\lambda=4$), and \emph{SO} adaptive headways. Notice logarithmic scale~\cite{Gershenson:2011a}.
}
\label{fig:homo}
\end{figure}
Results for a non-homogeneous scenario, with non equidistant stations nor initial positions of vehicles and unequal passenger demand ($\lambda$) at stations, are shown in Figure \ref{fig:nonhomo}. The default method collapses as well. The maximum method is not able to recover from the unequal initial headways and maintains them, leading also to high waiting times, even at stations, although not as high as for the default method. The self-organizing method is able to adapt to the non-homogeneous demands in this scenario and delivers a performance similar to that of the homogeneous scenario.
\begin{figure}[!ht]
\begin{center}
\includegraphics[width=3.5in]{img/results2}
\end{center}
\caption{
Results for non-homogeneous scenario. A. Passenger delays for methods: ``default" (\emph{DF}), ``max" (\emph{MX}), and ``self-organizing" (\emph{SO}), for different passenger inflow intervals $\lambda$. Lower boxes, slightly shifted to the right, at each column show waiting times at stations. Higher boxes show total waiting times. B. Headway standard deviations. Lower $\sigma_f$ implies more regular headways. Notice logarithmic scale~\cite{Gershenson:2011a}.
}
\label{fig:nonhomo}
\end{figure}
The self-organizing method is better than the theoretical optimum because of a slower-is-faster effect~\cite{Helbing:2000,Helbing:2009}. Passengers indeed wait more time at stations, but trying only to minimize passenger waiting time by forcing equal headways leads to \emph{friction} between vehicles, since vehicles serving stations with different passenger demands will idle and or leave passengers unattended at stations. Passenger inflow is not predictable, and assuming average flows to force predefined schedules will also lead to friction because of the same reason. On the contrary, the self-organizing method promotes \emph{synergy} by stigmergy of the vehicles, since they can balance---communicating through the antipheromones---the load of the system without idling and without collapsing, adapting to the current passenger demand at every station and state of the vehicles. These positive interactions allow the reduction of travel times, which benefit vehicles and passengers.
Traditional public transport regulation is more like clockwork, attempting to impose equal headways to changing demands. The self-organizing method is more like a healthy heart, where different intervals adapt to the instant demands of the system. Our current public transportation systems are more like diseased hearts: either too regular (cannot adapt) or arrhythmic (inefficient).
As this case study showed, living technology (adaptive, robust, self-organizing) can deliver a higher efficiency than that of traditional systems. Solutions to urban problems require the properties of living systems because problems are constantly changing. This limits their predictability and thus solutions which are unable to adapt to unforeseen situations. Since living systems make a living out of adapting to unforeseen situations, living technology is an excellent candidate for solving urban problems.
\section{Beyond the Metaphor: Towards Living Cities}
Cities will offer a higher quality of life if they exhibit the properties of livings systems.
After listing several current and potential urban living technologies, one can ask to what extent speaking about living cities is a mere metaphor and to what extent cities are usefully described as living systems.
Living systems are constantly adapting, learning and evolving because their environment is always changing at different timescales. Living systems also require to be robust to endure unforeseen perturbations. Efficient cities have to do the same. It is not enough being ``smart". The demands and conditions of cities change constantly at different scales, so they must adapt, learn and evolve in a robust fashion in order to endure. Cities are not physically similar to living systems (no DNA, no membranes), but functionally, they should exhibit the same properties. From a materialist point of view, it makes no sense to speak about living cities. However, from a functionalist point of view, it is very relevant to speak about the relationships between living systems, artificial life, living technology, and urban systems. This is because the properties of living systems (natural or artificial) can be exploited to solve urban problems, making cities more adaptive and robust.
If a notion of life based on entropy or information is used~\cite{Adami:1998,Gershenson:2007}, then one can even \emph{measure} to what extent different cities can be considered to be alive, with a continuous transition between non-living and living systems~\cite{Bedau1998}. In non-technical terms, if a city has a sufficient control over its own production, endowing it with a certain autonomy and integrity, then it can be usefully described as a living system. Living technology has been contributing to the increase of the ``liveness" of cities, as it was shown by the examples presented in this paper. Moreover, the study of ``living cities" is related to at least one of the open problems in artificial life~\cite{BedauEtAl2000}: To determine whether fundamentally novel living organizations can exist.
Technology has always evolved~\cite{Kelly:2010}, but with the aid of humans for most of its history. As living technology is developed, technology will be able not only to be more adaptive and robust, but to evolve by itself in directions that we cannot foresee. What can be said is that the integration between technology and living systems---including humans---will increase. Living cities will be the outcome of this integration.
Will solutions to urban problems using living technology bring new problems? Since predictability is limited, most probably new problems will arise. Nevertheless, we can always transform problems into opportunities. How? By deciding to do something about them.
\section*{Acknowledgments}
I should like to thank Steen Rasmussen and two anonymous referees for useful comments.
This work was partially supported by UNAM-DGAPA-IACOD project T2100111, by Intel\textsuperscript{\textregistered}, and SNI membership 47907 of CONACyT, Mexico.
\bibliographystyle{alj}
|
\section{Introduction}
By a convex body $S$ in the $m$-dimensional Euclidean space ${\mathbb R}^m$, we mean a compact and convex set $S\subset{\mathbb R}^m$
with non-empty interior, which we assume to be $0$-symmetric, i.e.\ $S=-S$.
A lattice $\Lambda$ is a discrete ${\mathbb Z}$-submodule of ${\mathbb R}^m$ of full rank.
Given a convex body $S$ and a lattice $\Lambda$ in ${\mathbb R}^m$,
the $i$-th successive minimum $\lambda_i(S,\Lambda)$ for $1\leq i\leq m$
of $S$ with respect to $\Lambda$ is defined as
\[\lambda_i(S,\Lambda)\coloneqq\inf\Set{\lambda>0 | \lambda S\cap\Lambda \text{ contains at least }i\text{ linearly independent elements}}\,.\]
With a convex body $S$ and a lattice $\Lambda$ we can associate the polar body
\[S^\star\coloneqq\Set{x\in{\mathbb R}^m | \left<x,y\right> \leq 1\ \forall\,y\in S}\]
and the polar lattice
\[\Lambda^\star\coloneqq\Set{x\in{\mathbb R}^m | \left<x,y\right> \in{\mathbb Z}\ \forall\,y\in \Lambda}\,,\]
where $\left<\,\cdot\,,\,\cdot\,\right>$ denotes the standard scalar product on ${\mathbb R}^m$.
We have $({\mathbb Z}^m)^\star={\mathbb Z}^m$ and $B_m^\star=B_m^{\phantom{\star}}$ for the Euclidean unit ball.
A classical inequality, first investigated by Mahler, is the transference result
\begin{equation}\label{eq:classicalinequality}
1\leq \lambda_i(S,\Lambda) \lambda_{m-i+1}(S^\star,\Lambda^\star) \leq m^{3/2}\,,
\end{equation}
for $1\leq i\leq m$. For the lower bound see Gruber \cite[\textsection\,5]{MR1242995},
while the upper bound follows from Banaszczyk \cite[Thm.~2.1]{banaszcyknewbounds}.
\medskip
We provide a generalisation of this inequality and of the notion of polarity
to the geometry of numbers
over the ring of adeles of an algebraic number field.
The theory of adelic geometry of numbers arises in the context of Siegel's Lemma,
which asks for a small integral solution to a system of linear equations with integer coefficients.
Answers by Thue, Siegel and others usually involve counting arguments or Minkowski's theorems on successive minima, cf \cite{schmidt.lnm1467}.
In order to allow coefficients and solutions from an algebraic number field,
Bombieri and Vaaler in \cite{bombierivaalersiegelslemma} proved
an adelic variant of Minkowski's second theorem on successive minima.
A comprehensive overview of adelic geometry of numbers can be found in \cite{thunderremarksonadelic}.
The theory has been further generalised, as has Siegel's Lemma, with recent results by
Fukshansky \cite{fukshanskysiegelslemma}, \cite{fukshanskyalgebraicpoints} and Gaudron \cite{gaudron}
on the number of algebraic points in bounded regions.
Further work on Siegel's Lemma for the algebraic closure of ${\mathbb Q}$
by Roy and Thunder~\cite{roythunderabsolutesiegel} involves the study of twisted heights.
Using these heights they introduce a different notion of adelic polarity
and an analogous statement of (\ref{eq:classicalinequality}) in terms of these heights,
which has most recently been extended by Rothlisberger~\cite{rothlisberger}.
The present paper however uses a more geometric approach,
directly extending the classical notion of polarity to the adelic setting.
To this end we fix an algebraic number field $K$ of degree $d$ over ${\mathbb Q}$,
with field discriminant $\Delta_K$, cf.~\cite[Kap.\,I]{neukirch}.
We will use geometry of numbers over the ring of adeles $K_{\mathbb A}$ of $K$ of rank $n\in{\mathbb N}$.
The definitions of an adelic convex body $S$
and the adelic successive minima $\lambda_i(S)$ for $1\leq i,j \leq n$,
will be provided in Section~\ref{sec:adelicgeometryofnumbers}.
For our definition of polar adelic body see Definition~\ref{def:adelicpolarbody}.
The main results of this paper are the following.
\begin{thm}\label{thm:adelicpolarupper}
Let $S$ be an adelic convex body, $S^\star$ its polar and let $\lambda_i(S)$,
$\lambda_j(S^\star)$ ($1\leq i,j\leq n$) be the successive minima of $S$ and $S^\star$ respectively.
Then for $1\leq\ell\leq n$
\[\lambda_\ell(S)\lambda_{n-\ell+1}(S^\star)\leq (nd)^{3/2} \,.\]
\end{thm}
In view of the classical result (\ref{eq:classicalinequality}) we are also interested in a lower bound,
which however we can not proof in full generality.
For a special class of adelic convex body and for $K$ totally real or a CM-field
(i.e.\ a field of complex multiplication) we get the following estimate.
\begin{thm}\label{thm:adelicpolarlower}
Let $K$ be totally real or a CM-field and
let $S$ be an adelic convex body, with the additional requirement that
for all complex places $v$, we have $S_v=\alpha S_v$ for $\alpha\in{\mathbb C}$ with $\abs{\alpha}=1$.
Let $S^\star$ be its polar and let $\lambda_i(S)$,
$\lambda_j(S^\star)$ ($1\leq i,j\leq n$) be the successive minima of $S$ and $S^\star$ respectively,
Then for $1\leq\ell\leq n$
\[\tfrac{1}{\sqrt[d]{\abs{\Delta_K}}}\leq \lambda_\ell(S)\lambda_{n-\ell+1}(S^\star)\,.\]
\end{thm}
Notice, that in the case $K={\mathbb Q}$ these results reduce to the classical statement (\ref{eq:classicalinequality}).
Finally, Example~\ref{exmp:bbQsqrt2} shows that the lower bound is sharp, at least for $n=1$.
\section{Adelic Geometry of Numbers}\label{sec:adelicgeometryofnumbers}
We start by giving a brief overview of the ring of adeles of an
algebraic number field $K$ of degree $d$ over ${\mathbb Q}$.
For more details and proofs we refer to \cite[Ch.~IV]{weilbasicnumber} and \cite[Ch.~VI]{knappadvancedalgebra}.
Let $r$ be the number of real and $s$ the number of pairs of complex embeddings
of $K$ into ${\mathbb C}$. Then $d=r+2s$. Denote by
${\mathcal O}$ the ring of algebraic integers of $K$ and by $\Delta_K$ its field discriminant.
Let $M(K)$ be its set of places.
For $v\in M(K)$ we write $v\nmid\infty$ for non-archimedian places
and $v\mid\infty$ for the archimedian ones.
For the corresponding absolute value on $K$ we write $\abs{\,\cdot\,}_v$.
We normalize it to extend either the usual absolute value on ${\mathbb Q}$ for archimedian places
or the $p$-adic absolute value for a prime $p$.
Then the local field $K_v$ is the completion of $K$ with respect to $v$.
For $v\nmid\infty$ let ${\mathcal O}_v$ be the local ring of integers.
Let $K_{\mathbb A}$ be the ring of adeles of $K$ and $K_{\mathbb A}^n$ the standard module of rank $n\geq 2$,
i.e.\ the $n$-fold product of adeles.
Recall that $K_{\mathbb A}$ is the restricted direct product of the $K_v$ with respect to the ${\mathcal O}_v$.
For any $v\in M(K)$ let $d_v=[K_v:{\mathbb Q}_v]$ be the local degree (${\mathbb Q}_\infty\mathrel{\cong}{\mathbb R}$).
Then for all primes $p\in{\mathbb Z}$
\begin{equation}\label{eq:productformulaetc}
d=\sum_{v\mid p} d_v\quad \text{and}\quad d=\sum_{v\mid \infty} d_v\,,\quad\text{and also}\quad
\prod_{v\in M(K)} \abs{a}_v^{d_v}=1
\end{equation}
for all non-zero $a\in K$.
Denote by $\sigma_i$, $1\leq i \leq r$ the embeddings of $K$ into ${\mathbb R}$
and by $\sigma_{r+i}=\overline{\sigma}_{r+i+s}$, $1\leq i\leq s$ the pairs
of embeddings of $K$ into ${\mathbb C}$, so $d=r+2s$.
We call $K$ \emph{totally real}, if $s=0$,
and we call $K$ a \emph{CM-field}, if it is a quadratic extension of a totally real field with $r=0$.
Then
\begin{align*}
\iota &\colon x\mapsto\bigl(\sigma_1(x),\ldots,\sigma_r(x),
\sigma_{r+1}(x),\ldots,\sigma_{r+s}(x)\bigr)\\
\shortintertext{and}
\overline{\iota} &\colon x\mapsto\bigl(\sigma_1(x),\ldots,\sigma_r(x),
\overline{\sigma}_{r+1}(x),\ldots,\overline{\sigma}_{r+s}(x)\bigr)
\end{align*}
are embeddings of $K$ into $K_\infty\coloneqq\prod_{v\mid\infty}K_v$.
There is a canonical isomorphism $\rho\colon K_\infty\rightarrow {\mathbb R}^{d}$ with
\[\begin{multlined}
\rho\bigl(x_1,\ldots,x_r,x_{r+1},\ldots,x_{r+s}\bigr)= \\
\qquad\bigl(x_1,\ldots,x_r,{\ensuremath{\operatorname{\frakR}}}(x_{r+1}),{\ensuremath{\operatorname{\frakI}}}(x_{r+1}),\ldots,{\ensuremath{\operatorname{\frakR}}}(x_{r+s}),{\ensuremath{\operatorname{\frakI}}}(x_{r+s})\bigr)\,.
\end{multlined}\]
Here ${\ensuremath{\operatorname{\frakR}}}$ and ${\ensuremath{\operatorname{\frakI}}}$ denote real and imaginary parts respectively.
Together we get $(\rho\circ\iota)\colon K\hookrightarrow {\mathbb R}^d$,
\[x\mapsto\bigl(\sigma_1(x),\ldots,\sigma_r(x),
{\ensuremath{\operatorname{\frakR}}}(\sigma_{r+1}(x)),{\ensuremath{\operatorname{\frakI}}}(\sigma_{r+1}(x)),\ldots,{\ensuremath{\operatorname{\frakR}}}(\sigma_{r+s}(x)),{\ensuremath{\operatorname{\frakI}}}(\sigma_{r+s}(x))\bigr)\,.\]
In the rank-$n$-case let $K_\infty^n\coloneqq\prod_{v\mid\infty}K_v^n$,
\[
\iota^n \coloneqq (\sigma_1^n,\ldots,\sigma_r^n,\sigma_{r+1}^n,\ldots,\sigma_{r+s}^n)
\colon K^n\rightarrow K_\infty^n\,,
\quad\text{$\overline{\iota}^n$ respectively,}\]
where the $\sigma_i$ act componentwise.
Similarily $\rho^n\colon K_\infty^n\rightarrow{\mathbb R}^{nd}$.
\begin{defn}\label{def:adelicconvexbody}
For each $v\nmid\infty$ let $S_v$ be a free ${\mathcal O}_v$-module of full rank,
where $S_v={\mathcal O}_v^n$ for all but finitely many $v$.
In other words, for any $v\nmid\infty$ there is an $A_v\in{\ensuremath{\operatorname{GL}}}_n(K_v)$ such that
$S_v=A_v^{-1}{\mathcal O}_v^n$, where $A_v$ is the identity for all but finitely many $v$.
For $v\mid\infty$ we have $K_v\mathrel{\cong}{\mathbb R}$ or $K_v\mathrel{\cong}{\mathbb C}$.
In this case let $S_v$ be a $0$-symmetric compact convex body with
non-empty interior in ${\mathbb R}^n$ or ${\mathbb C}^n\mathrel{\cong}{\mathbb R}^{2n}$ respectively.
Then the set
\[S = \prod_{v\nmid\infty} S_v \times \prod_{v\mid\infty} S_v\]
is called a closed symmetric \emph{adelic convex body}. If necessary, we denote $S_\infty=\prod_{v\mid\infty} S_v$.
\end{defn}
For $(x_v)_v\in K_{\mathbb A}^n$ we define the scalar multiple $(y_v)_v=\lambda(x_v)_v$ for $\lambda\in{\mathbb R}^+$ by
\[
y_v\coloneqq\begin{cases}\phantom{\lambda}x_v &\text{if } v\nmid\infty\,,\\
\lambda x_v &\text{if } v\mid\infty\,.\end{cases}
\]
\begin{defn}\label{def:adelsuccmindilat}
The \emph{$i$-th successive minimum} of the adelic convex body $S$ is
\[\lambda_i(S)=\inf\{\lambda>0 \mid \exists\, x_1,\ldots,x_i\in K^n\text{ lin.\ indep.\ over }K
\text{ s.t. } x_j\in\lambda S \text{ for all }j\}\]
for $1\leq i\leq n$. By construction $\lambda_i (S) \leq \lambda_j(S)$ for $i\leq j$.
\end{defn}
\begin{defn}\label{def:adelinhommin}
The \emph{inhomogeneous minimum} of the adelic convex body $S$ is
\[\mu(S)\coloneqq\inf\Bigl\{\mu>0 \Bigm| K_{\mathbb A}^n \subseteq \bigcup_{\zeta\in K^n} \bigl(\mu S+\zeta\bigr)\Bigr\}\,.\]
By construction $\mu(S) = \widehat{\mu}(\rho(S_\infty),\rho(\iota({\mathfrak M})))$,
where
\[\widehat{\mu}(T,\Lambda)\coloneqq\inf\Bigl\{\mu>0 \Bigm| {\mathbb R}^m\subseteq \bigcup_{\zeta\in \Lambda} \bigl(\mu T+\zeta\bigr) \Bigr\}\]
is the classical inhomogeneous minimum of the convex body $T\subset{\mathbb R}^m$ with respect to the lattice $\Lambda\subset{\mathbb R}^m$, cf.\ \cite[\textsection\,5]{MR1242995}.
Here ${\mathfrak M}=\bigcap_{v\nmid\infty} \bigl(S_v \cap K^n\bigr)$.
\end{defn}
\section{Adelic Polarity}
In order to define our notion of adelic polarity we first recall some
background from Algebraic Number Theory.
It is well-known~\cite[Ch.\,I,(2.8)]{neukirch}, that
\[T(x,y)\coloneqq{\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(xy)\]
is a non-degenerate symmetric ${\mathbb Q}$-bilinear form on $K$.
Here ${\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}$ denotes the field trace.
This allows to define
\begin{equation}
\prescript{\star}{}{\calO}\coloneqq\Set{x\in K | {\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(xy)\in{\mathbb Z}\ \forall\,y\in{\mathcal O}}\,,
\label{eq:defcompmod}
\end{equation}
the \emph{complementary module}, cf.~\cite[Ch.\,III,\,\textsection\,2]{neukirch}.
This is a fraction ideal in $K$, its inverse is the \emph{different} ${\mathfrak d}$.
On $K^n$ we get a bilinear form given by
\[T_n( x,y)\coloneqq\sum_{i=1}^n {\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(x_i y_i)\,.\]
\bigskip
By \cite[V\,§\,2,\,Thms.\,2\,\&\,3]{weilbasicnumber} for any fractional ideal ${\mathfrak m}$
there is a map $a\colon M(K)\rightarrow{\mathbb Z}$, such that ${\mathfrak m}$ can be written as
\begin{equation}\label{eq:idealasintersection}
{\mathfrak m}=\bigcap_{v\nmid\infty}(K\cap {\mathfrak p}_v^{a(v)})\,,
\end{equation}
where almost all $a(v)=0$ and ${\mathfrak p}_v$ is the unique maximal ideal in ${\mathcal O}_v$.
More concretely, we get the following special case.
\begin{lem}\label{lem:caloastalsschnitt}
Let $v\nmid\infty$ and define as in the global case
\[\prescript{\star}{}{\calO}_v\coloneqq\Set{x\in K_v | {\ensuremath{\operatorname{Tr}}}_{K_v/{\mathbb Q}_v}(xy)\in{\mathbb Z}_v\ \forall\,y\in{\mathcal O}_v}\,.\]
Then $\prescript{\star}{}{\calO}=\bigcap_{v\nmid\infty}(\prescript{\star}{}{\calO}_v\cap K)$.
For all but finitely many $v\nmid\infty$ we have $\prescript{\star}{}{\calO}_v={\mathcal O}_v$.
\end{lem}
\begin{proof}
By their definitions (cf.~\cite[p.\,377\,($\star$)]{knappadvancedalgebra}) we have
\[ \prescript{\star}{}{\calO}_v \cap K = \prescript{\star}{}{\calO}_{(v)} \coloneqq
\Set{ \tfrac{a}{b} | a \in \prescript{\star}{}{\calO}, b \in {\mathcal O} \setminus (v)}\supseteq\prescript{\star}{}{\calO},\]
where $\prescript{\star}{}{\calO}_{(v)}$ is the localisation of $\prescript{\star}{}{\calO}$ at the ideal $(v)$ corresponding to $v$.
For the converse inclusion we follow an idea by J.\,Jahnel\footnote{personal communication}.
Let $M\coloneqq \bigcap_{v\nmid\infty} \prescript{\star}{}{\calO}_{(v)} $, $x\in M$ and
consider the “ideal of denominators”
\[I\coloneqq\Set{ b\in {\mathcal O} | bx \in M}\,.\]
Since $x\in K\cap\prescript{\star}{}{\calO}_v=\prescript{\star}{}{\calO}_{(v)}$, we have $I\not\subset(v)$,
for the ideal in $K$ corresponding to $v$.
Since this holds for all $v$, we have $I={\mathcal O}$. Therefore $x\in\prescript{\star}{}{\calO}$.
The final statement follows from \cite[Lemma~6.48]{knappadvancedalgebra},
since only finitely many primes are ramified in $K$.
\end{proof}
We extend the construction from \eqref{eq:defcompmod} in a natural way to the rank-$n$ case with the form $T_n$.
\begin{lem}\label{lem:dimensionn}
Let $A\in{\ensuremath{\operatorname{GL}}}_n(K)$ and $A_v\in{\ensuremath{\operatorname{GL}}}_n(K_v)$ for any finite $v$.
Then
\[\prescript{\star}{}{(A{\mathcal O}^n)} = A^{-t} (\prescript{\star}{}{\calO})^n\quad\text{and}\quad
\prescript{\star}{}{(A_v^{\phantom{x}}{\mathcal O}_v^n)} = A_v^{-t} (\prescript{\star}{}{\calO}_v)^n\,.\]
\end{lem}
\begin{proof}
Notice, that
\[\prescript{\star}{}{({\mathcal O}^n)}\coloneqq\set{x \in K^n | T_n( x,y)\in{\mathbb Z}\ \forall\, y \in {\mathcal O}^n} \supseteq (\prescript{\star}{}{\calO})^n\,.\]
Suppose they are not the same, i.e.\ $\exists\,a\in\prescript{\star}{}{({\mathcal O}^n)}\setminus (\prescript{\star}{}{\calO})^n$.
Then for some $i$: $a_i\not\in\prescript{\star}{}{\calO}$, so there is some $b_i\in{\mathcal O}$,
such that ${\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(a_ib_i)\not\in{\mathbb Z}$ by definition of $\prescript{\star}{}{\calO}$.
But then $T_n(a,(0,\ldots,0,b_i,0,\ldots,0))\not\in{\mathbb Z}$ giving a contradiction.
Now let $(a_{ij})_{ij}=A\in{\ensuremath{\operatorname{GL}}}_n(K)$, $x,y\in K^n$.
Then
\begin{align*}
T_n(x, Ay)&=\sum_i {\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(x_i (Ay)_i)
=\sum_i {\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}\bigl(x_i \bigl({\textstyle\sum_j} a_{ij}y_j\bigr)\bigr)\\
&=\sum_i \sum_j{\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}\bigl(x_i ( a_{ij}y_j)\bigr)
=\sum_j \sum_i{\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}\bigl((a_{ij}x_i) y_j\bigr)\\
&=\sum_j {\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}((A^t x)_j y_j)=T_n( A^t x, y)\,.
\end{align*}
The second statement is obvious, as the above argument
works for $x,y\in K_v^n$ and $A_v\in{\ensuremath{\operatorname{GL}}}_n(K_v)$ verbatim using ${\ensuremath{\operatorname{Tr}}}_{K_v/{\mathbb Q}_v}$.
\end{proof}
\medskip
On the other hand, we can define a scalar product on ${\mathbb R}^d={\mathbb R}^{r+2s}$ as
\begin{equation}
(x,y)=\sum_{i=1}^r x_i y_i + 2 \sum_{i=r+1}^{2s} x_iy_i\,,\label{eq:skalprodaufRn}
\end{equation}
cf.~\cite[Ch.\,I,(5.1)]{neukirch}. This gives the scalar product
\begin{equation}\label{eq:skalprodaufKinfinity}
(x,y)\coloneqq(\rho(x),\rho(y))=\sum_{v\text{ real}} x_v y_v + \sum_{v\text{ complex}} (x_v\overline{y}_v +\overline{x}_v y_v)
\end{equation}
on $K_\infty$, cf.~\cite[p.\,222]{neukirch}.
\begin{lem}\label{lem:bilformsareequal}
For all $x,y\in K$
\[
{\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(x y) = \left( \rho(\iota(x)),\rho(\overline{\iota}(y))\right)\,.
\]
\end{lem}
\begin{proof}
Let $x,y\in K$, then
\begin{align*}
&\phantom{=}\left(\rho(\iota(x)),\rho(\overline{\iota}(y))\right)\\
&\phantom{=}=\sum_{j=1}^r \sigma_j(x) \sigma_j(y)
+ \sum_{j=1}^s 2\bigl({\ensuremath{\operatorname{\frakR}}}(\sigma_{r+j}(x)){\ensuremath{\operatorname{\frakR}}}(\overline{\sigma}_{r+j}(y))+ {\ensuremath{\operatorname{\frakI}}}(\sigma_{r+j}(x)){\ensuremath{\operatorname{\frakI}}}(\overline{\sigma}_{r+j}(y))\bigr)\\
&\phantom{=}=\sum_{j=1}^r \sigma_j(x) \sigma_j(y)
+ 2\sum_{j=1}^s \bigl({\ensuremath{\operatorname{\frakR}}}(\sigma_{r+j}(x)){\ensuremath{\operatorname{\frakR}}}(\sigma_{r+j}(y))- {\ensuremath{\operatorname{\frakI}}}(\sigma_{r+j}(x)){\ensuremath{\operatorname{\frakI}}}(\sigma_{r+j}(y))\bigr)\,.
\end{align*}
By~\cite[Ch.\,I,(2.6)\,(ii)]{neukirch}, ${\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(x)=\sum_\sigma \sigma(x)$,
where the sum is over all embeddings $\sigma\colon K\hookrightarrow \overline{{\mathbb Q}}$.
As all complex embeddings appear in conjugated pairs
\begin{align*}
{\ensuremath{\operatorname{Tr}}}_{K/{\mathbb Q}}(xy) &=
\sum_{j=1}^r \sigma_j(xy) + \sum_{j=1}^s \sigma_{r+j}(xy) + \sum_{j=1}^s \overline{\sigma}_{r+j}(xy) \\
&= \sum_{j=1}^r \sigma_{j}(x)\sigma_{j}(y) + 2\sum_{j=1}^s {\ensuremath{\operatorname{\frakR}}}(\sigma_{r+j}(x)\sigma_{r+j}(y))\,.
\end{align*}
The statement follows from ${\ensuremath{\operatorname{\frakR}}}(ab)={\ensuremath{\operatorname{\frakR}}}(a){\ensuremath{\operatorname{\frakR}}}(b)-{\ensuremath{\operatorname{\frakI}}}(a){\ensuremath{\operatorname{\frakI}}}(b)$.
\end{proof}
\begin{cor}\label{cor:polariscompl}
For any algebraic number field $K$ with ring of integers ${\mathcal O}$ and embeddings $\rho$ and $\iota$ as above, we have
\[\rho(\iota({\mathcal O}))^\star=\rho(\overline{\iota}(\prescript{\star}{}{\calO}))\,,\]
where $(\,\cdot\,)^\star$ is the polar with respect to the form in (\ref{eq:skalprodaufRn}).
\end{cor}
The scalar product $(\,\cdot\,,\,\cdot\,)$ on ${\mathbb R}^{nd}$ is also defined
as the sum of the components of each copy of ${\mathbb R}^d$.
Notice that we get the standard scalar product at the real places and
the real scalar product multiplied by $2$ at the complex places.
By direct consequence of Lemma~\ref{lem:dimensionn} and Corollary~\ref{cor:polariscompl}
and again \cite[V\,§\,2,\,Thm.\,2]{weilbasicnumber}, cf.\ (\ref{eq:idealasintersection}),
this leads to the following generalisation.
\begin{cor}\label{cor:scalarproductindimnd}
For any algebraic number field $K$ with ring of integers ${\mathcal O}$ and embeddings $\rho^n$ and $\iota^n$ as above, we have
\begin{align*}
\rho^n(\iota^n(A^{-1}{\mathcal O}^n))^\star&=\rho^n(\overline{\iota}^n(A^t(\prescript{\star}{}{\calO})^n))\\
\shortintertext{and}
\rho^n\Bigl(\iota^n\Bigl(\bigcap_{v\nmid\infty}(A_v^{-1}{\mathcal O}_v^n\cap K^n)\Bigr)\Bigr)^\star
&=\rho^n\Bigl(\overline{\iota}^n\Bigl(\bigcap_{v\nmid\infty}(A_v^t(\prescript{\star}{}{\calO}_v)^n\cap K^n)\Bigr)\Bigr)
\end{align*}
for $A\in{\ensuremath{\operatorname{GL}}}_n(K)$, $A_v\in{\ensuremath{\operatorname{GL}}}_n(K_v)$ for all $n\in{\mathbb N}$.
\end{cor}
\begin{rem}
Consider a finite number of $0$-symmetric convex bodies $S_i\subset{\mathbb R}^{m_i}$.
Then, using the classical notion of polarity,
\begin{equation}\label{eq:prodofkomplcontainscomplofprod}
\bigl(\prod_i S_i\bigr)^\star \subseteq \prod_i S_i^\star\,.
\end{equation}
Indeed, let $x\in\bigl(\prod_i S_i\bigr)^\star$, then $\left<x,y\right>\leq 1$
for all $y\in \prod_i S_i$. So especially for any $i$ we have
$\left<x,(0,\ldots,0,y_i,0,\ldots,0)\right>\leq 1$ for all $y_i\in S_i$.
But that implies $\left<x_i,y_i\right>\leq 1$ for all $i$,
which defines the right-hand side of (\ref{eq:prodofkomplcontainscomplofprod}).
For the scalar product $(\,\cdot\,,\,\cdot\,)$ instead of $\left\langle \,\cdot\,,\,\cdot\,\right\rangle$
we get
$\left<x,(0,\ldots,0,y_i,0,\ldots,0)\right>\leq \tfrac{1}{2}$ and
$\left<x_i,y_i\right>\leq \tfrac{1}{2}$ for the complex places ($x_i,y_i\in{\mathbb C}$),
so (\ref{eq:prodofkomplcontainscomplofprod}) holds as well.
\end{rem}
Due to Corollary~\ref{cor:scalarproductindimnd}, we are now in the situation to define our notion of adelic polarity.
\begin{defn}\label{def:adelicpolarbody}
Let $S = \prod_{v\nmid\infty} A_v^{-1}{\mathcal O}_v^n \times \prod_{v\mid\infty} S_v$ be an adelic convex body.
The \emph{polar adelic body} of $S$ is
\[S^\star\coloneqq\prod_{v\nmid\infty} A_v^{t} (\prescript{\star}{}{\calO}_v)^n \times\prod_{v\mid\infty} S_v^\star\,,\]
where $S_v^\star$ is the polar body of $S_v$ with respect to the restriction of (\ref{eq:skalprodaufRn}).
Since ${\mathcal O}_v=\prescript{\star}{}{\calO}_v$ for almost all $v\nmid\infty$ by Lemma~\ref{lem:caloastalsschnitt},
$S^\star$ is again an adelic convex body.
\end{defn}
\section{Adelic Transference Theorems}\label{sec:mainresults}
We now apply the results of the previous section,
especially Corollary~\ref{cor:scalarproductindimnd},
to prove the main results of this paper.
\begin{proof}[Proof of Theorem~\ref{thm:adelicpolarupper}]
Let
\[{\mathfrak M}=\bigcap_{v\nmid\infty} \bigl(A_v^{-1} {\mathcal O}_v^n \cap K^n\bigr)
\quad\text{and}\quad
{\mathfrak M}^\star=\bigcap_{v\nmid\infty} \bigl(A_v^{t} (\prescript{\star}{}{\calO}_v)^n \cap K^n\bigr)\,.\]
By \cite[Lemma]{thunderremarksonadelic} $\rho(\iota({\mathfrak M}))$
and $\rho(\overline{\iota}({\mathfrak M}^\star))$ are lattices of full rank in ${\mathbb R}^{nd}$.
By Corollary~\ref{cor:scalarproductindimnd}, they are polar to each other.
Denote by $S_\infty$ and $S_\infty^\star$ the infinite parts of
$S$ and $S^\star$ respectively.
By (\ref{eq:prodofkomplcontainscomplofprod}) we have
\begin{equation}\label{eq:dualindual}
(\rho(S_\infty))^\star\subset \rho(S_\infty^\star)\,.
\end{equation}
Denote by $\lambda_\ell(S)$ and $\lambda_\ell(S^\star)$ the adelic successive minima
of $S$ and $S^\star$ respectively
and by $\widehat{\lambda}_i(T,\Lambda)$ the classical successive minima
of the convex body $T$ and the lattice $\Lambda$ in ${\mathbb R}^{nd}$.
Then, by \cite[p.\,256]{thunderremarksonadelic}, for $\ell=1,\ldots,n$
\begin{align*}
\lambda_\ell(S) &\leq
\widehat{\lambda}_{(\ell-1)d+1}\bigl(\rho(S_\infty),\rho(\iota({\mathfrak M}))\bigr)\\
\shortintertext{and}
\lambda_\ell(S^\star) &\leq \widehat{\lambda}_{(\ell-1)d+1}
\bigl(\rho(S_\infty^\star),\rho(\overline{\iota}({\mathfrak M}^\star))\bigr)
\leq \widehat{\lambda}_{(\ell-1)d+1}
\bigl(\rho(S_\infty)^\star,\rho(\overline{\iota}({\mathfrak M}^\star))\bigr)\,,
\end{align*}
where the last inequality follows from (\ref{eq:dualindual}).
Finally, applying (\ref{eq:classicalinequality}), we conclude
\begin{align*}
\lambda_\ell(S)\lambda_{n-\ell+1}(S^\star)&\leq
\widehat{\lambda}_{(\ell-1)d+1}(\rho(S_\infty),\rho(\iota({\mathfrak M})))
\widehat{\lambda}_{((n-\ell+1)-1)d+1}(\rho(S_\infty)^\star,\rho(\overline{\iota}({\mathfrak M}^\star)))\\
&\leq \widehat{\lambda}_{(\ell-1)d+1}(\rho(S_\infty),\rho(\iota({\mathfrak M})))
\widehat{\lambda}_{(n-\ell)d+d}(\rho(S_\infty)^\star,\rho(\overline{\iota}({\mathfrak M}^\star)))\\
&\leq (nd)^{3/2} \,.\qedhere
\end{align*}
\end{proof}
\begin{cor}
Let $K$, $S$, $S^\star$ and $\lambda_1(S)$ be as in Theorem~\ref{thm:adelicpolarupper}
and let $\mu(S^\star)$ be the inhomogeneous minimum of $S^\star$. Then
\[\lambda_1(S) \cdot \mu(S^\star) \leq C\, nd(1+\log nd)\,,\]
where $C$ is a universal constant.
\end{cor}
\begin{proof}
As in the proof of Theorem~\ref{thm:adelicpolarupper},
we have $\lambda_1(S) = \widehat{\lambda}_1(\rho(\iota({\mathfrak M})),\rho(S_\infty))$
and by (\ref{eq:dualindual}) we get
\[\widehat{\mu}(\rho(S_\infty^\star),\Lambda) \leq \widehat{\mu}(\rho(S_\infty)^\star,\Lambda)\]
for any lattice $\Lambda\subset{\mathbb R}^{nd}$.
Therefore
\[\lambda_1(S) \cdot \mu(S^\star) \leq \widehat{\lambda}_1(\rho(S_\infty),\rho(\iota({\mathfrak M})))\cdot
\widehat{\mu}(\rho(S_\infty)^\star,\rho(\overline{\iota}({\mathfrak M}^\star))) \leq C\, nd(1+\log nd)\,,
\]
by \cite[Corollary~1]{banaszcykinequalities2} with some universal constant $C$.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{thm:adelicpolarlower}]
We use the standard bilinear form on $K^n$:
\[b(x,y)=\sum_{i=1}^n x_i \overline{y}_i\,,\]
where $\overline{\,\cdot\,}$ is complex conjugation if $K$ is a CM-field
and the identity for $K$ totally real.
Let $u_1,\ldots,u_n$ and $v_1,\ldots,v_n$ be $K$-bases of $K^n$ such that
$u_i\in \lambda_i(S) S$ and $v_j\in \lambda_j(S^\star) S^\star$ for all $i,j$.
Notice that for $u_i\in{\mathcal O}^n$ and $v_j\in(\prescript{\star}{}{\calO})^n$, we have
$b(A^{-1}u_j,\overline{A^{t}v_j})=b(u_j,\overline{v}_j)\in\prescript{\star}{}{\calO}$,
using that $\prescript{\star}{}{\calO}$ is a fractional ideal in $K$.
By definition of $\prescript{\star}{}{\calO}$ and the different ${\mathfrak d}$, we have
$\abs{x}\leq\abs{{\mathfrak d}}^{-1}$ for $x\in\prescript{\star}{}{\calO}$, \cite[III,\,2.1]{neukirch}.
This holds for any finite place $v$ as well.
Since $b$ is non-degenerate, there are
$i\in\Set{1,\ldots,\ell}$ and $j\in\Set{1,\ldots,n-\ell+1}$ such that $b(u_i,\overline{v}_j)\neq 0$.
Then by the product formula in (\ref{eq:productformulaetc})
\begin{align*}
1&=\prod_v \abs{b(u_i,\overline{v}_j)}_v^{d_v}\cdot
\left(\frac{\lambda_i(S)\lambda_j(S^\star)}{\lambda_i(S)\lambda_j(S^\star)}\right)^d\\
&=\prod_{v\nmid\infty} \bigl|b(u_i,\overline{v}_j)\bigr|_v^{d_v}
\cdot \left(\lambda_i(S)\lambda_j(S^\star)\right)^d
\cdot\prod_{v\mid\infty}
\bigl|b(\tfrac{1}{\lambda_i(S)}u_i,\tfrac{1}{\lambda_j(S^\star)}\overline{v}_j)\bigr|_v^{d_v}\,.
\end{align*}
Now for any finite $v$ we have $b(u_i,\overline{v}_j)\in\prescript{\star}{}{\calO}_v$, therefore $\bigl|b(u_i,\overline{v}_j)\bigr|_v^{d_v}\leq \abs{{\mathfrak d}_v}^{-d_v}$,
where ${\mathfrak d}_v$ denotes the local different.
Finally $\prod_{v\nmid\infty}\abs{{\mathfrak d}_v}^{-d_v}=\abs{\Delta_K}$, cf.~\cite[Ch.\,VI,\,§\,8]{knappadvancedalgebra}.
To conclude the proof, we consider the factors at the infinite places.
By assumption they are either all real or all complex. Fix some $v\mid\infty$.
Let $x\coloneqq\tfrac{1}{\lambda_i(S)}u_i$ and $y\coloneqq\tfrac{1}{\lambda_j(S^\star)}v_j$.
If $K$ is totally real, i.e.\ $v$ is real, we have
\[\bigl|b(x,\overline{y})\bigr|_v^{d_v}=\bigl|b(x,y)\bigr|_v^1
=\bigl|\sigma_v\bigl({\textstyle\sum}_i x_i y_i \bigr)\bigr|
=\bigl|{\textstyle\sum}_i \sigma_v(x_i) \sigma_v(y_i) \bigr|
\leq 1\,,\]
by definition of $S_v^\star$.
If $K$ is a CM-field, i.e.\ $v$ is complex, we get
\[\begin{multlined}
\bigl|b(x,\overline{y})\bigr|_v^{d_v}
=\bigl|\sigma_v\bigl({\textstyle\sum}_i x_i \overline{y}_i \bigr)\bigr|^2
=\bigl|{\textstyle\sum}_i \sigma_v(x_i) \overline{\sigma_v(y_i)} \bigr|^2\\
\leq \bigl(\bigl|{\ensuremath{\operatorname{\frakR}}}({\textstyle\sum}_i \sigma_v(x_i) \overline{\sigma_v(y_i)})\bigr|
+ \bigl|{\mathrm i}{\ensuremath{\operatorname{\frakI}}}({\textstyle\sum}_i \sigma_v(x_i) \overline{\sigma_v(y_i)})\bigr|\bigr)^2
\leq \left(\left|\tfrac{1}{2}\right|+1\left|\tfrac{1}{2}\right|\right)^2=1\,,\end{multlined}\]
by definition of $S_v^\star$, since ${\mathrm i}{\ensuremath{\operatorname{\frakI}}}(x)={\mathrm i}{\ensuremath{\operatorname{\frakR}}}({\mathrm i} x)$ for all $x\in{\mathbb C}$
and from $(\sigma_v(x_i))_i\in S_v$ we get ${\mathrm i}(\sigma_v(x_i))_i\in S_v$ by our additional requirement.
The conclusion follows from the monotonicity of the minima.
\end{proof}
\begin{exmp}\label{exmp:bbQsqrt2}
Let $n=1$ and $K={\mathbb Q}[\sqrt{2}]$, then ${\mathcal O}={\mathbb Z}[\sqrt{2}]={\mathbb Z}+\sqrt{2}{\mathbb Z}$
and the field discriminant is $\abs{\Delta_K}=8$.
Consider $x=a+b\sqrt{2}\in{\mathbb Q}[\sqrt{2}]$ and $y=c+d\sqrt{2}\in{\mathbb Z}[\sqrt{2}]$.
Then
\[xy=(a+b\sqrt{2})(c+d\sqrt{2})=ac+2bd+(ad+bc)\sqrt{2}\,.\]
Therefore
\[{\ensuremath{\operatorname{Tr}}}(xy)={\ensuremath{\operatorname{Tr}}}\begin{pmatrix}ac+2bd&2ad+2bc\\ ad+bc&ac+2bd\end{pmatrix}=2ac+4bd\]
and this is an integer if $a\in\tfrac{1}{2}{\mathbb Z}$ and $b\in\tfrac{1}{4}{\mathbb Z}$.
Therefore $\prescript{\star}{}{\calO}=\tfrac{1}{2}{\mathbb Z}+\tfrac{\sqrt{2}}{4}{\mathbb Z}$.
Now $\rho(\iota({\mathcal O})),\rho(\iota(\prescript{\star}{}{\calO}))\subset{\mathbb R}^2$ are lattices of rank $2$,
more precisely
\[\rho(\iota({\mathcal O}))=\begin{pmatrix}1&\sqrt{2}\\1&-\sqrt{2}\end{pmatrix}{\mathbb Z}^2\,,\qquad
\rho(\iota(\prescript{\star}{}{\calO}))=\begin{pmatrix}\tfrac{1}{2}&\tfrac{\sqrt{2}}{4}\\\tfrac{1}{2}&-\tfrac{\sqrt{2}}{4}\end{pmatrix}{\mathbb Z}^2\,,\]
and we see that $\rho(\iota({\mathcal O}))^\star=\rho(\iota(\prescript{\star}{}{\calO}))$.
This follows easily from the fact, that the matrices are the inverse transpose of one another.
Taking the $1$-dimensional unit ball $[-1,1]$ at both infinite places for the convex bodies,
we see that
\[S=\prod_{v\nmid\infty}{\mathcal O}_v\times\prod_{v\mid\infty}[-1,1]\quad\text{and}\quad
S^\star=\prod_{v\nmid\infty}\prescript{\star}{}{\calO}_v\times\prod_{v\mid\infty}[-1,1]
\]
are polar. Obviously $\lambda_1(S)\leq 1$ and since $\tfrac{\sqrt{2}}{4}<\tfrac{1}{2}$,
we have $\lambda_1(S^\star)\leq\tfrac{\sqrt{2}}{4}$.
This gives equality for the lower bound in Theorem~\ref{thm:adelicpolarlower}.
\end{exmp}
\medskip
\noindent{\itshape Acknowledgment.} I would like to thank Martin Henk,
Florian Heß, Matthias Henze, Jörg Jahnel and Kristin Stroth for helpful comments
and discussions on the subject.
\bibliographystyle{amsplain}
|
\section{Introduction}
We consider the following fractional $(q,p)$-Poincar\'e inequality in a bounded domain $G$ in $\varmathbb{R}^n\,,$
$n\geq 2\,,$
\begin{equation}\label{fractionalqp}
\int_G\vert u(x)-u_G\vert ^q\,dx
\le
c
\biggl(\int_G\int_{G\cap B^n(x,\tau \qopname\relax o{dist}(x,\partial G)
)}\frac{\vert u(x)-u(y)\vert ^p}{\vert x-y\vert ^{n+\delta p}}\,dy\,dx
\biggr)^{q/p}\,,
\end{equation}
where $1\le p,q <\infty$,
$\delta,\tau \in (0,1)$, and
the constant $c$ does not depend on $u\in L^p(G)$.
Our inequality \eqref{fractionalqp} with $q=p$ is stronger than
the fractional inequality
\begin{equation}\label{fractionalpp}
\int_G\vert u(x)-u_G\vert ^p\,dx
\le
c
\int_G\int_G\frac{\vert u(x)-u(y)\vert ^p}{\vert x-y\vert ^{n+\delta p}}\,dx\,dy\,,
\end{equation}
where on the right hand side is the commonly used seminorm on $W^{\delta,p}(G)$, \cite{A}.
Augusto C. Ponce showed that bounded Lipschitz domains support
the same type of inequalities
as \eqref{fractionalpp} but with general radial weights, \cite{P1}, \cite[Theorem 1.1]{P2}.
Jean Bourgain, Ha\"im Brezis, and Petru Mironescu
found the optimal constant $c$
in \eqref{fractionalpp} when $G$ is a cube \cite[Theorem 1]{BBM2}.
An elementary proof was provided by Vladimir Maz'ya and Tatyana Shaposhnikova, \cite{MS1},
\cite{MS2}.
The relationship between the right hand side of \eqref{fractionalpp} and the $L^p(G)$ integrability
of the absolute value of the gradient in smooth bounded domains is considered
in \cite{BBM1}.
We give sufficient geometric conditions for a bounded domain $G$ in $\varmathbb{R}^n$ to
support the fractional $(q,p)$-Poincar\'e inequality for $1\le q\le p<\infty$, Theorem \ref{thmP}. Examples of the domains which support the
fractional $(p,p)$-Poincar\'e inequality are
John domains, Theorem \ref{pp_John}. The John domains
include uniform domains and hence also Lipschitz domains.
We show that John domains
support the fractional Poincar\'e inequality
\eqref{fractionalqp} when $1<p\le q\le np/(n-\delta p)\,$ and $p<n/\delta\,,$ Theorem
\ref{fractional_sobolev_poincare}.
We also study more general bounded domains,
so called $s$-John domains with $s>1$. We prove
fractional $(1,p)$-Poincar\'e inequalities for
these domains, Theorem \ref{sharp}, and we show that these results are sharp, Theorem \ref{1p_counter}.
\section{Notation and auxiliary results}
We assume that $G$ is a bounded domain in Euclidean $n$-space
$\varmathbb{R}^n$, $n\geq 2$, throughout the paper.
We denote by
$\mathcal{D}$ the family of closed dyadic cubes in $\varmathbb{R}^n$. We let $\mathcal{D}_j$ be
the family of those dyadic cubes whose
side length is $2^{-j}$, $j\in\varmathbb{Z}$.
For a domain $G$ we fix its Whitney decomposition $W=W_G\subset\mathcal{D}$.
For the properties of dyadic cubes and Whitney cubes we refer to Elias M. Stein's book,
\cite{S}. We write $Q^*=\frac{9}{8}Q$ for $Q\in W$. Then,
\begin{equation}\label{dist_est}
\frac{3}{4}\qopname\relax o{diam}(Q)\le \qopname\relax o{dist}(x,\partial G)\le 6 \qopname\relax o{diam}(Q)\,,\quad \text{ if }x\in Q^*.
\end{equation}
Let us fix a cube $Q_0$ in the Whitney decomposition $W$.
For each $Q\in W$
there exists a chain of cubes
$(Q_0^*,Q_1^*,\cdots \,,Q_k^*)=:\mathcal{C}(Q^*)$
joining
two cubes $Q_0^*$ and $Q_k^*=Q^*$
such that
$
Q_i^*\cap Q_j^*\neq\emptyset\,
$
if and only if $\vert i-j\vert\le 1$.
The length of this chain is written as
$\ell(\mathcal{C}(Q^*)):=k$.
Once the chains of cubes have been
picked up, then for each Whitney cube $A$ we define a set
$A(W)=\{Q\in W\mid A^*\in \mathcal{C}(Q^*)\}$.
We call this construction a chain decomposition of $G$
with a fixed cube $Q_0$.
The side length of a cube $Q$ in $\varmathbb{R}^n$ is denoted by $\ell (Q)$.
We write $\chi_E$ for the characteristic function of a set $E$.
The Lebesque $n$-measure of a measurable set $E$ is denoted by $\vert E\vert.$
The upper Minkowski dimension of a set $E$ in $\varmathbb{R}^n$ is
\[
\dim_\mathcal{M}(E):=\sup \big\{\lambda\ge 0\mid \limsup_{r\to 0+} \mathcal{M}_\lambda(E,r)=\infty\big\},
\]
where
\[
\mathcal{M}_\lambda(E,r):=\frac{|E+B^n(0,r)|}{r^{n-\lambda}}=\frac{|\cup_{x\in E}B^n(x,r)|}{r^{n-\lambda}},\quad r> 0,
\]
is the $\lambda$-dimensional Minkowski precontent.
The notation $a\lesssim b$ is used to express that an estimate $a\le cb$ holds for some constant $c>0$
whose value is clear from the context.
We use subscripts to indicate the dependence on parameters, for example,
a quantity $c_{\lambda}$ depends on a parameter $\lambda$.
The following lemma gives a fractional inequality in a cube.
\begin{lemma}\label{inequality_cube}
Let $Q$ be a closed cube in $\varmathbb{R}^n$. Let $1\le q\le p <\infty $
and let $\delta,\rho \in (0,1)\,.$
Then, there is a constant $c<\infty $ independent of $u\in L^p(Q)$ such that
\begin{equation*}
\begin{split}
&\frac{1}{|Q|} \int_{Q} |u(y)-u_{Q}|^q\,dy
\\& \le c|Q|^{q(\delta /n-1/p)}\bigg(\int_{Q} \int_{Q\cap B^n(y,\,\rho \ell(Q))} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy\bigg)^{q/p}\,.
\end{split}
\end{equation*}
\end{lemma}
\begin{proof}
Without loss of generality we may assume that $Q=[0,1]^n$.
This comes from a simple scaling and translation argument.
Let us divide $Q$ into $k^n$ congruent and closed subcubes $Q_1,\ldots,Q_{k^n}$, where $k$ is chosen such that $R\subset B^n(y,\rho)$ for every $y\in R$ whenever
$R$ is a union of two cubes $Q_i$ and $Q_j$, $i,j\in \{1,2,\ldots,k^n\}$, sharing
a common face; in particular, the case $i=j$ is allowed.
We obtain
\begin{equation}\label{rest}
\begin{split}
\frac{1}{|R|} \int_{R} |u(y)-u_{R}|^q\,dy
&\le \bigg(\frac{1}{|R|} \int_{R} |u(y)-u_{R}|^p\,dy\bigg)^{q/p}\\
&\le \bigg(\frac{1}{|R|} \int_{R} \frac{1}{|R|} \int_{R} |u(y)-u(z)|^p\,dz\,dy\bigg)^{q/p}\\
&\lesssim |R|^{q(\delta/n-1/p)}\bigg(\int_{R} \int_{R} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy\bigg)^{q/p}\\
&\lesssim \bigg(\int_{Q} \int_{Q\cap B^n(y,\,\rho)} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy\bigg)^{q/p}\,.
\end{split}
\end{equation}
H\"older's inequality and Minkowski's inequality yield
\begin{align}\label{kaksi}
&\frac{1}{|Q|} \int_Q |u(y)-u_Q|^qdy \lesssim \frac{1}{|Q|} \int_Q |u(y)-u_{Q_1}|^qdy \notag\\
&\lesssim \sum_{j=1}^{k^n} \int_{Q_j}|u(y)-u_{Q_j}|^qdy + \sum_{j=1}^{k^n} \int_{Q_j} |u_{Q_j}-u_{Q_1}|^q dy\,.
\end{align}
By \eqref{rest} it is enough to estimate the second series in \eqref{kaksi}. Let us
fix $Q_j$, $j\in \{1,\ldots,k^n\}$, and let
$\sigma:\{1,2,\ldots,kn\}\to \{1,2,\ldots,k^n\}$ be
such that
$\sigma(1)=1$, $\sigma(kn)=j$, and the subsequent cubes
$Q_{\sigma(i)}$ and $Q_{\sigma(i+1)}$ share
a common face if $i=1,\ldots,kn-1$.
Since $kn\lesssim 1$, we obtain
\begin{align}\label{sums}
|u_{Q_{j}}-u_{Q_1}|^q &\le \bigg(\sum_{i=1}^{kn-1} |u_{Q_{\sigma(i+1)}}-u_{Q_{\sigma(i)}}|\bigg)^q\notag\\
&\lesssim \sum_{i=1}^{kn-1} |u_{Q_{\sigma(i+1)}}-u_{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}|^q+\sum_{i=1}^{kn-1}|u_{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}-u_{Q_{\sigma(i)}}|^q.
\end{align}
Let us consider the first sum in \eqref{sums}. Note that
\begin{align*}
&|u_{Q_{\sigma(i+1)}}-u_{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}|^q\\&\le \frac{1}{|Q_{\sigma(i+1)}|}\int_{Q_{\sigma(i+1)}} |u_{Q_{\sigma(i+1)}} -u(y) + u(y)-u_{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}|^q dy\\
&\lesssim \frac{1}{|Q_{\sigma(i+1)}|}\int_{Q_{\sigma(i+1)}} |u(y)-u_{Q_{\sigma(i+1)}}|^qdy\\
&\qquad + \frac{1}{|{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}|} \int_{{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}} |u(y)-u_{{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}}|^q dy.
\end{align*}
By \eqref{rest} we obtain
\[
\sum_{i=1}^{kn-1} |u_{Q_{\sigma(i+1)}}-u_{Q_{\sigma(i+1)}\cup Q_{\sigma(i)}}|^q
\lesssim \bigg(\int_{Q} \int_{Q\cap B^n(y,\,\rho)} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy\bigg)^{q/p}.
\]
Similar estimates for the remaining sum in \eqref{sums} conclude the proof.
\end{proof}
We also need some estimates involving porous sets in $\varmathbb{R}^n$.
\begin{definition}\label{porous}
A set $S$ in Euclidean $n$-space
is {\em porous in $\varmathbb{R}^n$} if for some $\kappa\in (0,1]$
the following statement is true:
for every $x\in\varmathbb{R}^n$ and $0<r\le 1$ there is $y\in B^n(x,r)$ such that
$B^n(y,\kappa\,r)\cap S=\emptyset$.
\end{definition}
The following lemma gives a norm estimate related to
porous sets, and it is based on maximal function techniques. This lemma might be of independent interest.
\begin{lemma}\label{por_lem}
Suppose that $S$ is porous in $\varmathbb{R}^n$ and let $1\le p<\infty$.
If
$x\in S$ and $0<r\le 1$, then
\[
\int_{B^n(x,r)} \log^p\frac{1}{\qopname\relax o{dist}(y,S)}\,dy
\le cr^n(1+\log^p r^{-1}),
\]
where the constant $c$ is independent of $x$ and $r$.
\end{lemma}
\begin{proof}
Let us write
\[
\mathcal{C}_{S} = \{R\in\mathcal{D}\,:\,\mathrm{dist}(x_R,S)/(4+\sqrt n)\le \ell(R)\le 1\},
\]
where $x_R$ is the midpoint of a dyadic cube $R$.
Suppose that
$R\in\mathcal{D}$ is such that $\ell(R)\le 1$ and
$\qopname\relax o{dist}(y,S)\le 4\ell(R)$ for some $y\in R$. Then, since
\begin{equation}\label{r_inclusion}
\begin{split}
\qopname\relax o{dist}(x_R,S)&\le \qopname\relax o{dist}(x_R,y) + \qopname\relax o{dist}(y,S)\\&\le \sqrt n\ell(R) + \qopname\relax o{dist}(y,S) \le (4+\sqrt n)\ell(R)
\end{split}
\end{equation}
for the midpoint of $R$, we conclude that $R\in\mathcal{C}_S$.
Fix $j\in \varmathbb{N}_0$ such that $2^{-j}\le r< 2^{-j+1}$, and
consider a dyadic cube $Q\in\mathcal{D}_{j}$ for which
$Q\cap B^n(x,r)\not=\emptyset$.
By covering $B^n(x,r)$ with
such dyadic cubes it is enough to show that
\begin{equation}\label{test}
|| \log\mathrm{dist}(\cdot,S)^{-1} ||_{L^p(Q\cap B^n(x,r))}^p
\lesssim r^n(1+\log^p r^{-1}).
\end{equation}
By the porosity and the Lebesgue density theorem, the $n$-measure
of $S$ is zero.
Hence, it is enough to consider points $y\in Q\cap B^n(x,r)\setminus S$. Since $x\in S$,
\begin{equation}\label{q_est}
1\le \frac{2\ell(Q)}{\qopname\relax o{dist}(y,S)}.
\end{equation}
Let us consider a finite sequence of dyadic cubes
\[
Q=Q_0(y)\supset Q_1(y)\supset \dotsb \supset Q_m(y),
\]
each of them containing the point $y$ and satisfying
\begin{equation}\label{ratio}
\ell(Q_i(y))/\ell(Q_{i+1}(y))=2,\qquad i=0,1,\ldots,m-1.
\end{equation} The
last cube is chosen to satisfy
\begin{equation}\label{qsat}
\qopname\relax o{dist}(y,S)/4 \le \ell(Q_m(y)) < \qopname\relax o{dist}(y,S)/2.
\end{equation}
From \eqref{q_est} it follows that $m\ge 1$.
By \eqref{ratio} and \eqref{q_est}
\[
2^m = \prod_{i=0}^{m-1} \frac{\ell(Q_i(y))}{\ell(Q_{i+1}(y))} =\frac{\ell(Q_0(y))}{\ell(Q_m(y))}
>\frac{2\ell(Q_0(y))}{\qopname\relax o{dist}(y,S)}=\frac{2\ell(Q)}{\qopname\relax o{dist}(y,S)}\ge 1.
\]
Hence,
\[
m\ge \log 2^m\ge \log 2\ell(Q) - \log \qopname\relax o{dist}(y,S)\ge 0.
\]
Furthermore, \eqref{qsat} and \eqref{r_inclusion} yield
$Q_i(y)\in \mathcal{C}_S$ if $i=0,1,\ldots,m$.
Thus, we obtain
\begin{align*}
\sum_{\substack{R\in\mathcal{C}_S\\varmathbb{R}\subset Q}} \chi_R(y) \ge 1+m &\ge 1+\log (\ell(Q)) - \log \qopname\relax o{dist}(y,S
\ge 0,
\end{align*}
where $\chi_R$ is the characteristic function of $R$.
Integrating this inequality and using triangle-inequality yields
\begin{align*}
&||\log \qopname\relax o{dist}(\cdot,S)^{-1} ||_{L^p(Q\cap B^n(x,r))}
\\&\le |1+\log \ell(Q)|\,|Q\cap B^n(x,r)|^{1/p} +\bigg\| \sum_{\substack{R\in\mathcal{C}_S\\varmathbb{R}\subset Q}} \chi_R \bigg\|_{L^p(\varmathbb{R}^n)}.
\end{align*}
Since $S$ is porous in $\varmathbb{R}^n$, we may
follow the proof of
\cite[Theorem 2.10]{ihna}. We obtain a constant
$K_\kappa$, depending on $\kappa$ in Definition \ref{porous}, and
families
\[\{\hat R\}_{R\in\mathcal{C}_S^k},\quad \mathcal{C}_S^k\subset \mathcal{C}_S,\qquad k=0,1,\ldots,K_\kappa-1,\]
where each $\{\hat R\}_{R\in\mathcal{C}_S^k}$ is a disjoint family of cubes $\hat R\subset R$, such that
\begin{align*}
\bigg\| \sum_{\substack{R\in\mathcal{C}_S\\varmathbb{R}\subset Q}} \chi_R \bigg\|_{L^p(\varmathbb{R}^n)}
\lesssim \sum_{k=0}^{K_\kappa-1} \bigg\| \sum_{\substack{R\in\mathcal{C}_S^k\\varmathbb{R}\subset Q}} \chi_{\hat R} \bigg\|_{L^p(\varmathbb{R}^n)}
\le \sum_{k=0}^{K_\kappa-1} ||\chi_Q||_{L^p(\varmathbb{R}^n)}\lesssim |Q|^{1/p}.
\end{align*}
By combining the estimates we obtain
\begin{align*}
||\log \qopname\relax o{dist}(\cdot,S)^{-1} ||_{L^p(Q\cap B^n(x,r))}
\lesssim (1+\log\ell(Q)^{-1})|Q|^{1/p}
\lesssim (1+\log r^{-1})r^{n/p}.
\end{align*}
Estimate \eqref{test} follows.
\end{proof}
\section{Conditions for the fractional Poincar\'e inequality}
In the following theorem we give sufficient conditions for
a bounded domain to support
the fractional $(q,p)$-Poincar\'e inequality \eqref{fractionalqp}.
\begin{theorem}\label{thmP}
Let $G$ be a bounded domain in $n$-dimensional Euclidean space, $n\geq 2\,,$
with a Whitney decomposition $W$.
Let $1\le q\le p<\infty\,$ and let $\delta,\tau\in (0,1)$.
(1) If $q<p$ and if there exists a chain decomposition of $G$ such that
\begin{equation}\label{sharpe}
\sum_{A\in W} \Bigg(\sum_{Q\in A(W)}
\ell(\mathcal{C}(Q^*))^{q-1}|Q|\, |A|^{q(\delta /n-1/p)}\Bigg)^{p/(p-q)}<\infty\,,
\end{equation}
then $G$ supports the fractional $(q,p)$-Poincar\'e inequality
\eqref{fractionalqp}.
(2) If $q=p$ and if there exists a chain decomposition of $G$ such that
\begin{equation}\label{pp}
\sup_{A\in W}
\sum_{Q\in A(W)}
\ell(\mathcal{C}(Q^*))^{p-1}|Q|\, |A|^{p\delta /n-1}<\infty\,,
\end{equation}
then $G$ supports the fractional $(p,p)$-Poincar\'e inequality
\eqref{fractionalqp}.
\end{theorem}
\begin{proof}
We prove (1); the proof of (2) is similar. Let $\delta$ and $\tau$ in $(0,1)$ be given.
We use H\"older's inequality and Minkowski's inequality and then the Whitney decomposition
to obtain
\begin{align}\label{integraltriangle}
\int_G \vert u(x)-u_G\vert^q\,dx
&\lesssim\int_{G}\vert u(x)-u_{Q_0^*}\vert^q\,dx\notag
\\&\le \sum_{Q\in W_{}}\int_{Q^*}
\vert u(x)-u_{Q_0^*}\vert ^q\,dx \notag\\
&\lesssim \sum_{Q\in W_{}}\int_{Q^*}\vert u(x)-u_{Q^*}\vert^q\,dx
+ \sum_{Q\in W_{}}\int_{Q^*}
\vert u_{Q^*}-u_{Q_0^*}\vert ^q\,dx\,.
\end{align}
Lemma \ref{inequality_cube} with $\rho=2\tau/3$ yields
\begin{align*}
&\int_{Q^*}
\vert u(x)-u_{ Q^*}\vert ^q\,dx
\\&\lesssim \vert Q^*\vert^{1+q(\delta /n -1/p)}\biggl(
\int_{ Q^*}\int_{Q^*\cap B^n(y,\,\rho \ell (Q^*))}
\frac{\vert u(z)-u(y)\vert^p}{\vert z-y\vert^{n+\delta p}}\,dz\,dy
\biggr)^{q/p}\,.
\end{align*}
Inequalities \eqref{dist_est} and $(1+q\delta /n-q/p )(p/(p-q))>1$
imply
\begin{align*}
&\sum_{Q\in W_{}}\int_{ Q^*}\vert u(x)-u_{ Q^*}\vert^q\,dx\\
&\lesssim \sum_{Q \in W} |Q|^{
1+q\delta /n-q/p}
\biggl(\int_{Q^*} \int_{Q^*\cap B^n(y,\,\rho \ell (Q^*))} \frac{|u(y)-u(z)|^p}{|z-y|^{n+{\delta}p}}\,dz\,dy\bigg)^{q/p}
\\
&\le \bigg( \sum_{Q\in W} (
|Q|^{1+q\delta /n-q/p} )^{p/(p-q)} \bigg)^{(p-q)/p}\\
&\qquad\qquad\qquad
\bigg( \sum_{Q\in W} \int_{Q^*}\int_{Q^*\cap B^n(y,\,\rho \ell (Q^*))} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy \bigg)^{q/p}\\
&\lesssim
\bigg( \int_{G} \int_{G\cap B^n(y,\tau \qopname\relax o{dist}(y,\partial G))} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy \bigg)^{q/p}.
\end{align*}
Next, we estimate the latter sum in (\ref{integraltriangle}).
By using chains from the chain decomposition we obtain
\begin{align*}
\sum_{Q\in W_{}}\int_{ Q^*}
\vert u_{ Q^*}-u_{ Q^*_0}\vert ^q\,dx
&\lesssim
\sum_{Q\in W}\vert Q\vert\,
\biggl(\sum_{j=1}^{k}
\vert u_{ Q^*_{j}}-
u_{ Q^*_{j-1}}
\vert \biggr)^q\\
&\le
\sum_{Q\in W}
{\ell(\mathcal{C}(Q^*))}^{q-1}\vert Q\vert\, \biggl(\sum_{j=1}^{k}
\vert u_{ Q^*_{j}}-
u_{ Q^*_{j-1}}
\vert^q\biggr)\,.
\end{align*}
Estimate $\max\{|Q^*_j|,|Q^*_{j-1}|\}\lesssim |Q^*_j\cap Q^*_{j-1}|$
and H\"older's inequality yield
\begin{align*}
\vert u_{ Q^*_{j}}-
u_{ Q^*_{j-1}}
\vert^q
&\lesssim
\sum_{i=j-1}^{j}
\biggl(\vert Q^*_{i}\vert ^{-1}\int_{ Q^*_i}
\vert u(x)-u_{ Q^*_i}\vert\,dx\biggr)^q\\
&\le
\sum_{i=j-1}^{j}
\vert Q^*_i\vert ^{-1}\int_{ Q^*_i}
\vert u(x)-u_{ Q^*_i}\vert^q\,dx\,.
\end{align*}
Lemma \ref{inequality_cube} with $\rho=2\tau/3$ implies
\begin{align*}
&\vert u_{ Q^*_{j}}-
u_{ Q^*_{j-1}}
\vert^q\\
&\lesssim
\sum_{i=j-1}^{j}
\vert Q^*_i\vert^{q(\delta /n -1/p)}
\biggl(
\int_{ Q^*_i}\int_{Q_i^*\cap B^n(y,\,\rho \ell (Q_i^*))}
\frac{\vert u(z)-u(y)\vert^p}{\vert z-y\vert^{n+\delta p}}\,dz\,dy\biggr)^{q/p}\,.
\end{align*}
We have obtained for the second sum in \eqref{integraltriangle}
\begin{align*}
&\sum_{Q\in W_{}}\int_{ Q^*}
\vert u_{ Q^*}-u_{ Q^*_0}\vert ^q\,dx\\
&\lesssim\sum_{Q\in W}
{\ell(\mathcal{C}(Q^*))}^{q-1}\vert Q\vert
\\&\qquad\qquad\biggl(
\sum_{j=0}^{k}
\vert Q^*_j\vert ^{q(\delta /n-1/p)}
\biggl(\int_{ Q^*_j}\int_{Q_j^*\cap B^n(y,\,\rho \ell (Q_j^*)}
\frac{\vert u(z)-u(y)\vert^p}{\vert z-y\vert^{n+\delta p}}\,dz\,dy
\bigg)^{q/p}\biggr)\,.
\end{align*}
When we rearrange the double sum, we obtain
\begin{align*}
&\sum_{Q\in W_{}}\int_{ Q^*}
\vert u_{ Q^*}-u_{ Q^*_0}\vert ^q\,dx\\
&\lesssim \sum_{A\in W} \sum_{Q\in A(W)}
\ell(\mathcal{C}(Q^*))^{q-1} |Q|\,|A|^{q(\delta /n-1/p)}\\
&\qquad\qquad\qquad
\bigg(\int_{A^*} \int_{A^*\cap B^n(y,\,\rho \ell (A^*))} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy\bigg)^{q/p}\,.
\end{align*}
H\"older's inequality with
$\left ( \frac{p}{q},\frac{p}{p-q}\right)$, and
inequalities (\ref{sharpe}) and \eqref{dist_est} yield
\begin{align*}
\sum_{Q\in W_{}}\int_{ Q^*}
\vert u_{ Q^*}-u_{ Q^*_0}\vert ^q\,dx
&\lesssim
\bigg(\sum_{A\in W} \int_{A^*} \int_{A^*\cap B^n(y,\,\rho \ell (A^*))} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy \bigg)^{q/p}\\
&\lesssim
\bigg(\int_{G}\int_{G\cap B^n(y,\tau\qopname\relax o{dist}(y,\partial G))} \frac{|u(y)-u(z)|^p}{|z-y|^{n+\delta p}}\,dz\,dy \bigg)^{q/p}.
\end{align*}
Hence, $G$ supports
the fractional $(q,p)$-Poincar\'e inequality \eqref{fractionalqp}.
\end{proof}
\begin{remark}
Let $G$ be a dounded domain in $\varmathbb{R}^n$ and let $1\le p<\infty\,.$
By \cite[Theorem 6.6]{Hu} the estimate
\begin{equation}\label{poincarecondition}
\sup_{A\in W}\sum_{Q\in A(W)}\ell(\mathcal{C}(Q^*))^{p-1}\,|Q|\,|A|^{p/n-1}
<\infty
\end{equation}
is a sufficient condition for the classical $(p,p)$-Poincar\'e inequality
to be valid in the domain $G$.
A comparison to our sufficient condition \eqref{pp}
for the {\em fractional} $(p,p)$-Poincar\'e inequality
shows that
condition \eqref{poincarecondition} for the classical $(p,p)$-Poincar\'e inequality is weaker.
\end{remark}
\section{Positive results for $1$-John domains}\label{pos_1_John}
As an application of Theorem \ref{thmP} we show
that $1$-John domains support the fractional
$(p,p)$-Poincar\'e inequality, Theorem \ref{pp_John}.
We also
consider fractional Sobolev--Poincar\'e inequalities,
Theorem \ref{fractional_sobolev_poincare} and Remark \ref{remaining_results}.
We recall that
bounded uniform and Lipschitz domains are examples of
$1$-John domains.
\begin{definition}\label{sjohn}
A bounded domain $G$ in $\varmathbb{R}^n$, $n\ge 2$, is an {\em $s$-John domain}, $s\ge 1$, if
there is a point $x_0$ in $G$ and a constant $c>0$ such that every point $x$ in $G$ can be joined
to $x_0$ by a rectifiable curve $\gamma:[0,l]\to G$ parametrized by its arc length
for which $\gamma(0)=x$, $\gamma(l)=x_0$, $l\le c$, and
\[
\qopname\relax o{dist}(\gamma(t),\partial G)\ge t^s/c\quad \text{ for }t\in [0,l].
\]
The point $x_0$ is called an {\em $s$-John center} of $G$.
\end{definition}
If $G$ is a $1$-John domain, then its boundary $\partial G$ is porous in $\varmathbb{R}^n$,
Definition \ref{porous}. The boundary of an $s$-John domain with $s>1$
may have positive Lebesgue $n$-measure, \cite{N1}, and thus it is not
necessarily porous in $\varmathbb{R}^n$.
Let us construct a chain decomposition
of a given $s$-John domain $G$.
Let $Q\in W=W_G$ and fix a rectifiable curve
$\gamma$ that is parametrized by its arc length
and joins the midpoints $x_Q$ and $x_0:=x_{Q_0}$, Definition \ref{sjohn}.
Assume that $x_{Q_0}$ lies in one of the
cubes intersecting $Q$. Join
$x_Q$ to $x_{Q_0}$ by an arc that is contained in $Q\cup Q_0$ and
whose length is comparable to $\ell(Q)$.
Otherwise there is $r>0$
such that $\gamma(r)$ lies
in the boundary of a Whitney cube $P$ that intersects $Q$ and
$\gamma(t)$ belongs to a cube
that is not intersecting $Q$ whenever $t\in (r,\ell(\gamma)]$.
Join the midpoint
$x_Q$ to the midpoint $x_P$ by an
arc whose length is comparable to $\ell(Q)$ and
is in $Q\cup P$.
We iterate these steps
with $Q$ replaced by $P$, and we continue
until we reach $x_{Q_0}$.
Let $\gamma_Q$ be this composed curve parametrized by its arc length.
It is straightforward
to verify that $\ell(\gamma_Q)\le c$ and
\begin{equation}\label{esg}
t^s/c \le \,\qopname\relax o{dist}(\gamma_Q(t),\partial G)\qquad \text{ if }\,t\in[0,\ell(\gamma_Q)],
\end{equation}
where $c>0$ depends
on the $s$-John constant of $G$, $s$, and $n$.
Let $\mathcal{C}(Q^*)$ be a chain consisting
of cubes $A^*$ such that
$A\in W$ and $x_A\in \gamma_Q[0,\ell(\gamma_Q)]$.
For $1$-John domains we first have the following result.
\begin{theorem}\label{pp_John}
A $1$-John domain $G$ in $\varmathbb{R}^n$
supports the fractional $(p,p)$-Poincar\'e inequality
\eqref{fractionalqp} if $1\le p<\infty$ and
$\tau,\delta\in(0,1)$.
\end{theorem}
\begin{proof}
We may assume
that $\qopname\relax o{diam}(G)\le 1$.
By \eqref{esg} with $s=1$ and the fact that $\gamma_Q$, $Q\in W$, connects
the midpoints of cubes in $\mathcal{C}(Q^*)$,
\begin{equation}\label{chain_length}
\ell(\mathcal{C}(Q^*))\le c\bigg(1+\log \frac{1}{\ell(Q)}\bigg),
\end{equation}
where the constant $c$ is independent of $Q$.
If $A\in W$, then
\begin{equation}\label{contains}
\bigcup_{Q\in A(W)} Q\subset B^n(\omega_A,\qopname\relax o{min}\{1,c\ell(A)\}),
\end{equation}
where
$\omega_A$ is the closest point in $\partial G$ to $x_A$ and
the constant $c>0$ is independent of $A$.
By \eqref{chain_length} and \eqref{contains} we obtain
\begin{align*}
&\sum_{Q\in A(W)}
\ell(\mathcal{C}(Q^*))^{p-1}|Q|
\lesssim
\sum_{Q\in A(W)}
|Q|\,\bigg(1+\log \frac{1}{\ell(Q)}\bigg)^{p-1}\\
&\lesssim \sum_{Q\in A(W)}
|Q|\,\bigg(1+\log^{p}\frac{1}{\ell(Q)}\bigg)\lesssim \sum_{Q\in A(W)}
\int_Q\bigg(1+\log^{p}\frac{1}{\qopname\relax o{dist}(y,\partial G)}\bigg)\,dy\\
&\le \int_{B^n(\omega_A,\qopname\relax o{min}\{1,c\ell(A)\})} \bigg(1+\log^{p}\frac{1}{\qopname\relax o{dist}(y,\partial G)}\bigg)\,dy.
\end{align*}
Since $\partial G$ is porous in $\varmathbb{R}^n$, Lemma \ref{por_lem} yields
\[
\sum_{Q\in A(W)}
\ell(\mathcal{C}(Q^*))^{p-1}|Q|
\lesssim
|A|(1+\log^p \ell(A)^{-1})\lesssim |A|^{1-\delta p/n}.
\]
We have verified condition \eqref{pp} in Theorem \ref{thmP}. Hence,
the domain $G$ supports the fractional $(p,p)$-Poincar\'e inequality.
\end{proof}
We state an immediate corollary of Theorem \ref{pp_John}.
\begin{corollary}
Let $G$ be a bounded domain in $\varmathbb{R}^n$, $n\ge 2$, and let $1\le p<\infty$,
$\delta,\tau \in (0,1)$.
Then $G$ supports the fractional $(p,p)$-Poincar\'e inequality \eqref{fractionalqp}
if $G$ is a uniform domain or a Lipschitz domain.
\end{corollary}
It is well known \cite[Theorem 5.1, Lemma 3.1]{B} that
$1$-John domains support Sobolev--Poincar\'e inequalities:
if $1\le p\le q\le np/(n-p)$, $p<n$, then there
is $c>0$ such that, for
every $u\in W^{1,p}(G)$,
\begin{equation}\label{sobolev_poincare}
\bigg(\int_G |u(x)-u_G|^q\,dx\bigg)^{1/q}\le c\bigg(\int_G |\nabla u(x)|^p\,dx\bigg)^{1/p}.
\end{equation}
We consider
the corresponding fractional Sobolev--Poincar\'e inequalities
on $1$-John domains, Theorem \ref{fractional_sobolev_poincare}.
For the proof of this theorem we need the Riesz potentials
$I_\delta$, $\delta\in (0,n)$, that are defined for suitable $f$ by
\[
I_\delta(f)(x) = \int_{\varmathbb{R}^n} \frac{f(y)}{|x-y|^{n-\delta}} dy.
\]
A proof of the following theorem is in \cite[Theorem 1]{He}.
\begin{theorem}\label{hedberg}
Let $0<\delta<n$, $1<p<q<\infty$, and $1/p-1/q=\delta/n$. Then
$||I_\delta(f)||_q \le c||f||_p$ for a constant $c>0$ independent
of $f\in L^p(\varmathbb{R}^n)$.
\end{theorem}
We also need the following chaining lemma. It is
a slight modification
of \cite[Theorem 9.3]{HK}:
we add the new condition 3 but the proof
adapts to our setting, and
we omit the details.
\begin{lemma}\label{chaining_lemma}
Let $G$ in $\varmathbb{R}^n$ be a $1$-John domain
whose $1$-John constant is $c_J>1$.
Fix a number $M>1$. Denote by $x_0\in G$ the $1$-John center of $G$, and let
\[
B_0:=B(x_0,\mathrm{dist}(x_0,\partial G)/4Mc_J).
\]
Then, there is a constant $c>0$, depending on $G$, $M$, and $n$,
as follows: given $x\in G$ there
is a sequence of balls
$B_i = B(x_i,r_i)\subset G$, $i=0,1,\ldots$, such that for all $i=0,1,\ldots$, the
following conditions {\em 1--5} hold:
\begin{itemize}
\item[1.] $|B_i\cup B_{i+1}|\le c|B_i\cap B_{i+1}|$;
\item[2.] $\mathrm{dist}(x,B_i)\le cr_i$;
\item[3.] $\mathrm{dist}(B_i,\partial G)\ge Mr_i$;
\item [4.] $|x-x_i|\le cr_i$ and $r_j\to 0$ as $j\to\infty$;
\item [5.] $\sum_{j=0}^\infty \chi_{B_j}\le c\chi_G$.
\end{itemize}
\end{lemma}
The following result is a fractional Sobolev--Poincar\'e inequality
for $1$-John domains.
\begin{theorem}\label{fractional_sobolev_poincare}
Assume that $G$ is a $1$-John domain in $\varmathbb{R}^n$, $n\ge 2$.
Suppose that $\tau,\delta\in (0,1)$, $p<n/\delta$, and
\[
1<p\le q\le \frac{np}{n-\delta p}.
\]
Then $G$ supports
the fractional $(q,p)$-Poincar\'e inequality \eqref{fractionalqp}.
\end{theorem}
\begin{proof}
By H\"older's inequality we may assume that $q=np/(n-\delta p)$.
Fix $\tau\in (0,1)$ and let $u\in L^p(G)$.
Let $x\in G$ be a Lebesgue point of $u$, and consider the
associated balls $B_i=B(x_i,r_i)$ from Lemma \ref{chaining_lemma}
satisfying conditions 1--5 with $M>2/\tau$.
The following holds: for all $i$,
\begin{equation}\label{ball_incl}
B_i\subset B^n(y,\tau\qopname\relax o{dist} (y,\partial G)),\qquad \text{ if }y\in B_i.
\end{equation}
Namely,
let us fix $y\in B_i$ and let $z$ be any point in $B_i\,$.
Then, by condition 3 in Lemma \ref{chaining_lemma},
\begin{align*}
\vert z-y\vert
&\le \vert y-x_i\vert+\vert x_i-z\vert
\le 2r_i\le 2\frac{\qopname\relax o{dist} (B_i,\partial G)}{M}\\
&\le \frac{2}{M}\qopname\relax o{dist} (y,\partial G) <\tau\qopname\relax o{dist} (y,\partial G).
\end{align*}
By the Lebesgue differentiation theorem and condition 4 in Lemma \ref{chaining_lemma},
\begin{equation*}
u(x)=\lim_{i\to\infty}\frac{1}{\vert B_i\vert}\int_{B_i}u(y)\,dy=\lim_{i\to\infty} u_{B_i}\,.
\end{equation*}
Hence, by condition 1 in Lemma \ref{chaining_lemma}, we obtain
\begin{align*}
\vert u(x)-u_{B_0}\vert
&\le\sum_{i=0}^{\infty}\vert u_{B_i}-u_{B_{i+1}}\vert\\
&\le\sum_{i=0}^{\infty}\Bigl(\vert u_{B_i}-u_{B_i\cap B_{i+1}}\vert
+\vert u_{B_{i+1}}-u_{B_i\cap B_{i+1}}\vert\Bigr)\\
&\lesssim\sum_{i=0}^{\infty}
\frac{1}{|B_i|}\int_{B_i}\vert u(y)-u_{B_i}\vert\,dy\,.
\end{align*}
For a ball $B_i$,
\begin{equation}\label{ball_est}
\begin{split}
\frac{1}{\vert B_i\vert}\int_{B_i}\vert u(y)-u_{B_i}\vert\,dy
&=\frac{1}{\vert B_i\vert}\int_{B_i}
\bigg\vert
\frac{1}{\vert B_i\vert}
\int_{B_i} ( u(y)-u(z))\,dz\bigg\vert\,dy\\
&\le \frac{1}{\vert B_i\vert}\int_{B_i}
\biggl(
\frac{1}{\vert B_i\vert}
\int_{B_i}\vert u(y)-u(z)\vert^p\,dz\biggr)^{1/p}\,dy\\
&=\frac{1}{\vert B_i\vert^{1+1/p}}
\int_{B_i}\biggl(
\int_{B_i}\vert u(y)-u(z)\vert^p\,dz\biggr)^{1/p}\,dy\\
&\lesssim
\vert B_i\vert^{\delta /n -1}\int_{B_i}\biggl(\int_{B_i}
\frac{\vert u(y)-u(z)\vert^p}{\vert y-z\vert^{n+\delta p}}\,dz\biggr)^{1/p}\,dy\,.
\end{split}
\end{equation}
Let us write
\begin{equation*}
g(y):=\biggl(\int_{G\cap B^n(y,\tau\qopname\relax o{dist}(y,\partial G))}\frac{\vert u(y)-u(z)\vert ^p}{\vert y-z\vert ^{n+\delta p}}\,dz\biggr)^{1/p}\,.
\end{equation*}
By \eqref{ball_est}, \eqref{ball_incl} and condition 2 in Lemma \ref{chaining_lemma},
\begin{align*}
\sum_{i=0}^{\infty}\frac{1}{\vert B_i\vert}
\int_{B_i}\vert u(y)-u_{B_i}\vert \,dy
&\lesssim
\sum_{i=0}^{\infty}\vert B_i\vert^{\delta /n-1}\int_{B_i}
\biggl(\int_{B_i}\frac{\vert u(y)-u(z)\vert ^p}{\vert y-z\vert ^{n+\delta p}}\,dz\biggr)^{1/p}\,dy
\\
&\le
\sum_{i=0}^{\infty}\vert B_i\vert^{\delta /n-1}\int_{B_i}
\biggl(\int_{B^n(y,\tau\qopname\relax o{dist}(y,\partial G))}\frac{\vert u(y)-u(z)\vert ^p}{\vert y-z\vert ^{n+\delta p}}\,dz\biggr)^{1/p}\,dy\\
&\lesssim
\sum_{i=0}^{\infty}r_i^{n(\delta /n-1)}\int_{B_i}g(y)\,dy\\
&\lesssim
\sum_{i=0}^{\infty}\int_{B_i}\frac{g(y)}{\vert x-y\vert ^{n-\delta}}\,dy\,.
\end{align*}
By condition 5 in Lemma \ref{chaining_lemma},
\begin{equation}
\vert u(x)-u_{B_0}\vert\lesssim
\int_{G}\frac{g(y)}{\vert x-y\vert ^{n-\delta}}\,dy=I_{\delta}(\chi_G g)(x)
\end{equation}
for every Lebesgue point $x\in G$.
By integrating this inequality and using
Theorem \ref{hedberg}, we obtain
\begin{align*}
\biggl(\int_G \vert u(x)-u_{B_0}\vert ^{q}\,dx\biggr)^{1/q}
&\lesssim
\|I_\delta(\chi_G g)\|_q \lesssim \|\chi_G g\|_p\\
&=
\biggl(\int_{G}\int_{G\cap B^n(y,\tau\,\qopname\relax o{dist}(y,\partial G))}\frac{\vert u(y)-u(z)\vert ^p}{\vert y-z\vert ^{n+\delta p}}\,dz\,dy\biggr)^{1/p}\,.
\end{align*}
Inequality \eqref{fractionalqp} follows.
\end{proof}
\begin{remark}\label{remaining_results}
The proof
of Theorem \ref{fractional_sobolev_poincare}
also gives the following result:
Suppose that $G$ is a $1$-John domain in $\varmathbb{R}^n$. Let
$\tau,\delta\in (0,1)$ and let $p,q\in [1,\infty)$ be such that
\[
0\le 1/p-1/q<\delta/n.
\]
Then $G$ supports
the fractional $(q,p)$-Poincar\'e inequality \eqref{fractionalqp}.
Indeed, it suffices to recall that the linear operator $f\mapsto I_\delta(\chi_G f)$
is bounded from $L^p(G)$ to $L^q(G)$,
\cite[Lemma 7.12]{GT}.
\end{remark}
\section{Positive results for $s$-John domains with $s>1$}
We prove
the fractional $(1,p)$-Poincar\'e inequality
\eqref{fractionalqp} for $s$-John domains,
Theorem \ref{sharp}. We show in Section \ref{sharp_sec}
that this result is sharp in terms of the restriction on $p$,
Theorem \ref{1p_counter}.
\begin{theorem}\label{sharp} Let
$s>1$, $1<p<\infty$, $\lambda \in [n-1,n)$, and let $\delta,\tau\in (0,1)$.
Suppose that
\begin{equation}\label{oletus}
s<\frac{n+1-\lambda}{1-\delta},\quad p>\frac{s(n-1)-\lambda+1}{n-s(1-\delta)-\lambda+1}.
\end{equation}
Let $G$ be an $s$-John domain in $\varmathbb{R}^n$
such that $\mathrm{dim}_{\mathcal{M}}(\partial G)\le \lambda$.
Then $G$ supports the fractional $(1,p)$-Poincar\'e inequality \eqref{fractionalqp}.
\end{theorem}
We need preparations for the proof of Theorem \ref{sharp}.
By scaling we may assume that $\qopname\relax o{diam}(G)\le 1$. Hence,
the side lengths of all Whitney cubes in $W=W_G$ are bounded by one
and
\begin{equation}\label{w_identity}
W=\bigcup_{j=0}^\infty W_j,
\end{equation}
where each
$W_j$ stands for the family of cubes $A\in W$ with
$\ell(A)=2^{-j}$.
For a given $s$-John domain $G$, we
consider its chain decomposition that is constructed in Section \ref{pos_1_John}.
Given $j,k\in\varmathbb{N}$ and $\sigma\ge 1$ we define
\[
W_{j,k,\sigma}:=
\{A\in W_j\mid 2^{-(j-k)n}\le |\cup A(W)\,|\le \sigma\cdot 2^{-(j-k-1)n}\}.
\]
The following
lemma from \cite[Lemma 4.7]{HH-SV} gives the properties we need for
this chain decomposition of $G$.
The integer part of $\alpha\in\varmathbb{R}$ is denoted by $[\alpha]$.
\begin{lemma}\label{sest}
Let $s>1$ and let $G$ be an $s$-John domain in $\varmathbb{R}^n$
such that $\qopname\relax o{diam}(G)\le 1$ and $\mathrm{dim}_{\mathcal{M}}(\partial G)< \lambda\in [n-1,n)$.
Then, there is a constant $\sigma\ge 1$ such that
\begin{equation}\label{kirjoitus}
W_j= \bigcup_{k=0}^{[j-j/s]} W_{j,k,\sigma}\qquad \text{ for every } j\in \varmathbb{N}.
\end{equation}
Furthermore, if $k\in \{0,1,\ldots,[j-j/s]\}$, then
\begin{equation}\label{sid}
\sharp W_{j,k,\sigma}\le c2^{-kn} 2^{j(n+1+(\lambda-n-1)/s)}.
\end{equation}
The positive constant $c$ depends on $n$, $s$, $\partial G$, and the $s$-John constant
of the domain $G$.
\end{lemma}
We are ready for the proof of Theorem \ref{sharp}.
\begin{proof}[{\it Proof of Theorem \ref{sharp}}]
Choose $\lambda'\in (\lambda,n)$ such that
\eqref{oletus} is true if $\lambda$ is replaced
by $\lambda'$. Then
$\mathrm{dim}_{\mathcal{M}}(\partial G)<\lambda'$ and hence we may assume
that $\mathrm{dim}_{\mathcal{M}}(\partial G)$ is
strictly less than $\lambda\in [n-1,n)$.
By Theorem \ref{thmP}
it is enough to prove the finiteness of
\[
\Sigma := \sum_{A\in W} \bigg(\sum_{Q\in A({W})}
|Q|\, |A|^{\delta/n-1/p}\bigg)^{p/(p-1)}
=\sum_{A\in W}\big( |\cup A(W)\,|\, |A|^{\delta/n-1/p} \big)^{p/(p-1)},
\]
where the chain decomposition of $G$ is given by Lemma \ref{sest}.
By \eqref{w_identity} and \eqref{kirjoitus} in Lemma \ref{sest}
\begin{align*}
\Sigma &=\sum_{j=0}^\infty
\sum_{k=0}^{[j-j/s]}
\sum_{A\in W_{j,k,\sigma}} \big( |\cup A(W)\,|\, |A|^{\delta/n-1/p}\big)^{p/(p-1)}.
\end{align*}
Then, by using the definition of $ W_{j,k,\sigma}$ and \eqref{sid}
from Lemma \ref{sest} we obtain
the estimate
\begin{align*}
\Sigma&\lesssim \sum_{j=0}^\infty \sum_{k=0}^{[j-j/s]} 2^{-kn} 2^{j(n+1+(\lambda-n-1)/s)}\cdot \big( 2^{-(j-k)n}\cdot 2^{-jn(\delta/n-1/p)}\big)^{p/(p-1)}\\
&= \sum_{j=0}^\infty \sum_{k=0}^{[j-j/s]} 2^{kn(p/(p-1)-1)} 2^{j(n+1+(\lambda-n-1)/s-np/(p-1)-\delta p/(p-1)+n/(p-1))}.
\end{align*}
Let us fix $j$ and $k$ as in the summation above. Then,
\[
kn\bigg(\frac{p}{p-1}-1\bigg) \le n(j-j/s)\bigg(\frac{p}{p-1}-1\bigg)
=\frac{jn(1-1/s)}{p-1}.
\]
The trivial estimate $[j-j/s]\le j$ implies that
\begin{align*}
\Sigma &\lesssim \sum_{j=0}^\infty j\cdot 2^{j(n(1-1/s)/(p-1)+n+1+(\lambda-n-1)/s-np/(p-1)-\delta p/(p-1)+n/(p-1))}\\
&=\sum_{j=0}^\infty j\cdot 2^{j(ns-s+\lambda p-\lambda-np-p+1-p(\delta-1)s)/s(p-1)}.
\end{align*}
By \eqref{oletus}
the last series converges.
\end{proof}
\section{Sharpness of Theorem \ref{sharp}}\label{sharp_sec}
We show
that Theorem \ref{sharp} is sharp
by proving Theorem \ref{1p_counter}. For this purpose we construct $s$-John domains which do not support
the fractional $(1,p)$-Poincar\'e inequality \eqref{fractionalqp}
for certain values of $p$.
Let us recall the construction of the $s$-version of a given $1$-John domain $G$,
\cite{HH-SV}.
We
may assume that the diameter of $G$ is restricted by condition
\begin{equation}\label{w_restr}
(\ell(Q)/8)^s \le \ell(Q)/32,\qquad \text{ if }Q\in W_G.
\end{equation}
Let $Q$ be a closed cube in $\varmathbb{R}^n$ centered at $x=(x_1,\ldots,x_n)$ and
whose side length $\ell=\ell(Q)$ satisfies
$(\ell/8)^s\le \ell/32$.
Thus,
$Q=\prod_{i=1}^n\,[x_i-\ell/2,x_i+\ell/2]$.
The {\em room} in $Q$ is the open cube
\[
R(Q):=\mathrm{int}(\frac{1}{4}Q)=\prod_{i=1}^{n} (x_i-\ell/8,x_i+\ell/8)
\]
centered at $x$ with side length $\ell/4$.
The {\em $s$-passage} in $Q$ is the open set
\[
P_s(Q):= \bigg(\prod_{i=1}^{n-1} \big(x_i-(\ell/8)^s,x_i+(\ell/8)^s \big)\bigg) \times
(x_n+\ell/8,x_n+\ell/4).
\]
Since
$(\ell/8)^s < \ell/8$, we have $P_s(Q)\subset \frac{1}{2}Q$.
The {\em long $s$-passage} in $Q$ is the open set
\[
L_s(Q):= \bigg(\prod_{i=1}^{n-1} \big(x_i-(\ell/8)^s,x_i+(\ell/8)^s \big)\bigg) \times
(x_n,x_n+\ell/2)\subset Q.
\]
The {\em $s$-apartment} in $Q$ is the set
\begin{equation}\label{sap}
A_s(Q):= L_s(Q)\cup (Q\setminus (\partial R(Q)\cup \partial P_s(Q)))\subset Q.
\end{equation}
\begin{definition}\label{s_version}
Let $G$ in $\varmathbb{R}^n$ be a $1$-John domain and let $s>1$ be
a number such that \eqref{w_restr} holds.
Then,
the {\em s-version of $G$} is the domain
\[
G_s := Q_0\cup \bigcup_{\substack{Q\in W_G\\Q\not=Q_0}} A_s(Q).
\]
Here
$Q_0\in W_G$ is the cube containing the $1$-John center $x_0$ of $G$.
\end{definition}
We
construct test functions. Let $Q\in W_G$ be fixed, and
define
the {\em tiny $s$-passage} in $Q$ to be the open set
\[
T_s(Q):=
\bigg(\prod_{i=1}^{n-1} \big(x_i-(\ell/8)^s,x_i+(\ell/8)^s \big)\bigg) \times
(x_n+5\ell/32,x_n+7\ell/32).
\]
Then, we define a continuous function
\[u^{A_s(Q)}\colon G_s\to \varmathbb{R}\]
which has linear decay along the $n^{\text{\tiny th}}$ variable in $T_s(Q)$
and
is constant in both components of $P_s(Q)\setminus T_s(Q)$, and satisfies
\begin{equation}\label{values}
u^{A_s(Q)}(x) =\begin{cases}
\ell(Q)^{(\lambda-n)/q},\qquad &\text{ if }x\in R(Q);\\
0,\qquad &\text{ if } x\in G_s\setminus (R(Q)\cup P_s(Q)).
\end{cases}
\end{equation}
In the sense of distributions in $G_s$,
\begin{equation}\label{gradient}
\nabla u^{A_s(Q)} = (0,\ldots,0,-16\ell(Q)^{(\lambda-n)/q-1}\chi_{T_s(Q)})
\end{equation}
pointwise almost everywhere.
The reason why we do not let $u^{A_s(Q)}$ have
linear decay along the whole $s$-passage $P_s(Q)$ is that
we need the following property.
\begin{remark}\label{loc_rem}
Let $Q\in W_G$. Suppose that
$x\in G_s$ and $y\in B^n(x,\qopname\relax o{dist}(x,\partial G_s))$ are such that
\[|u^{A_s(Q)}(x)-u^{A_s(Q)}(y)|\not=0.\] Then
$x$ and $y$ both belong to $P_s(Q)$.
This fact follows from the
assumption \eqref{w_restr}.
\end{remark}
The following proposition is the main tool for
proving Theorem \ref{1p_counter}.
\begin{prop}\label{sharp_counter}
Let $G$ be a $1$-John domain in $\varmathbb{R}^n$ and $s>1$
be
such that \eqref{w_restr} holds. Suppose that
\[
\limsup_{k\to \infty} 2^{-\lambda k}\cdot\sharp W_k
>0,\qquad \text{ where }\lambda=\mathrm{dim}_{\mathcal{M}}(\partial G)\in [n-1,n).
\]
Let
$\delta,\tau\in (0,1)$ and
$1\le q<p<\infty$ be such that
\begin{equation}\label{rels}
\frac{(p-q)(\lambda-n)}{pq} + \frac{(s-1)(n-1)}{p} \ge 1-s(1-\delta).
\end{equation}
Then
the $s$-version of $G$
is an $s$-John domain
with $\mathrm{dim}_{\mathcal{M}}(\partial G_s)=\lambda$
and $G_s$ does not support
the fractional $(q,p)$-Poincar\'e inequality \eqref{fractionalqp}.
\end{prop}
\begin{proof}
The fact
\[\mathrm{dim}_{\mathcal{M}}(\partial G_s)=\mathrm{dim}_{\mathcal{M}}(\partial G)=\lambda\]
is from \cite[Proposition 5.11]{HH-SV}.
By \cite[Proposition 5.16]{HH-SV}, the domain
$G_s$ is an $s$-John domain.
Hence, it remains
to prove the failure of the fractional Poincar\'e inequality.
Let us choose
$k_0\in\varmathbb{N}$ such that
\begin{equation*}\label{number}
\limsup_{k\to \infty} 2^{-\lambda(k-k_0)}\cdot \sharp W_k>2.
\end{equation*}
This allows us to
choose indices
$j(k)$, $k\in \varmathbb{N}$, inductively such that
\[\max\{k_0,-\log_2\ell(Q_0)\}< j(1)<j(2)<\dotsb\]
and
$\sharp W_{j(k)} \ge 2\cdot 2^{\lambda (j(k)-k_0)}$ for every $k\in \varmathbb{N}$.
Let us write $M_j:=2^{[\lambda(j-k_0)]}$, where $[\lambda(j-k_0)]$ means the integer part
of $\lambda(j-k_0)$, and let us
choose cubes \[Q_{j(k)}^1,\ldots,Q_{j(k)}^{2M_{j(k)}}\in W_{j(k)}\setminus \{Q_0\}.\]
For every $m\in\varmathbb{N}$ we define
\[
v_m := \sum_{k=1}^m \bigg(\sum_{i=1}^{M_{j(k)}} u^{A_s(Q_{j(k)}^i)}-\sum_{i=M_{j(k)}+1}^{2M_{j(k)}} u^{A_s(Q_{j(k)}^i)}\bigg).
\]
Note that
$(v_m)_{G_s}=0$ and
\begin{align*}
A_m:&=\bigg(\int_{G_s} |v_m - (v_m)_{G_s}|^q\bigg)^{1/q}
= \bigg(\sum_{k=1}^m\sum_{i=1}^{2M_{j(k)}} \int_{G_s}|u^{A_s(Q_{j(k)}^i)}|^q\bigg)^{1/q}
\\&\ge
\Big(m\cdot2\cdot 2^{\lambda(j(k)-k_0)-1}\cdot
2^{-j(k)(\lambda-n)}
\cdot
4^{-n}\cdot 2^{-j(k)n}\Big)^{1/q} =
c_{n,q,\lambda,k_0}m^{1/q}.
\end{align*}
Next we estimate the right hand side of \eqref{fractionalqp} with $u=v_m$.
We write
\[
G_s(x):=B^n(x,\qopname\relax o{dist}(x,\partial G_s))\subset G_s\qquad \text{ for } x\in G_s.
\]
Remark \ref{loc_rem} yields: if
$x\in G_s$ and $y\in G_s(x)$ are such that
$|v_m(x)-v_m(y)|\not=0$, then
$x,y\in P_s(Q)$ for some Whitney cube $Q\in W_G$.
By using this we obtain
\begin{align*}
B_m:&=\bigg(\int_{G_s} \int_{G_s(x)} \frac{|v_m(x)-v_m(y)|^p}{|x-y|^{n+\delta p}}\,dy\,dx\bigg)^{1/p}\\
&=\bigg(\sum_{Q\in W_G} \int_{Q\cap G_s}\int_{G_s(x)}\frac{|v_m(x)-v_m(y)|^p}{|x-y|^{n+\delta p}}\,dy\,dx\bigg)^{1/p}\\
&=\bigg(\sum_{Q\in W_G} \int_{P_s(Q)}\int_{P_s(Q)\cap G_s(x)}\frac{|v_m(x)-v_m(y)|^p}{|x-y|^{n+\delta p}}\,dy\,dx\bigg)^{1/p}\\
&=\bigg(\sum_{k=1}^m\sum_{i=1}^{2M_{j(k)}} \int_{P_s(Q_{j(k)}^i)}\int_{P_s(Q_{j(k)}^i)\cap G_s(x)}\frac{|u^{A_s(Q_{j(k)}^i)}(x)-u^{A_s(Q_{j(k)}^i)}(y)|^p}{|x-y|^{n+\delta p}}\,dy\,dx\bigg)^{1/p}.
\end{align*}
Let us fix a cube $R=Q_{j(k)}^i$, where
$k\in \{1,\ldots,m\}$ and $i\in \{1,2,\ldots,2M_{j(k)}\}$.
By \eqref{gradient}
\[
|u^{A_s(R)}(x)-u^{A_s(R)}(y)|\le 16\ell(R)^{(\lambda-n)/q-1}|x-y|,\qquad x,y\in P_s(R).
\]
Hence,
\begin{align*}
\mathcal{I}_R:=&\int_{P_s(R)}\int_{P_s(R)\cap G_s(x)}\frac{|u^{A_s(R)}(x)-u^{A_s(R)}(y)|^p}{|x-y|^{n+\delta p}}\,dy\,dx\\
&\lesssim \ell(R)^{p(\lambda-n)/q-p}\int_{P_s(R)}\int_{P_s(R)\cap G_s(x)} |x-y|^{-n+(1-\delta)p}\,dy\,dx.
\end{align*}
Note that $G_s(x)\subset B^n(x,\ell(R)^s)$ if $x\in P_s(R)$. Thus,
\[
\int_{P_s(R)\cap G_s(x)} |x-y|^{-n+(1-\delta)p}dy\le \int_{B^n(0,\ell(R)^s)} |y|^{-n+(1-\delta)p}dy
\lesssim \ell(R)^{s(1-\delta)p},
\]
and it follows that
\begin{align*}
\mathcal{I}_R&\lesssim \ell(R)^{p(\lambda-n)/q-p}|P_s(R)|\ell(R)^{s(1-\delta)p}\\&= \ell(R)^{p(\lambda-n)/q-p+s(n-1)+1+s(1-\delta)p}=2^{-j(k)(p(\lambda-n)/q-p+s(n-1)+1+s(1-\delta)p)}.
\end{align*}
These estimates and inequality \eqref{rels} yield
\begin{align*}
B_m&\lesssim \bigg(\sum_{k=1}^m
2^{\lambda j(k)}2^{-j(k)(p(\lambda-n)/q-p+s(n-1)+1+s(1-\delta)p)}\bigg)^{1/p}\lesssim m^{1/p}.
\end{align*}
By using the assumption $q<p$ we obtain
\[
\frac{A_m}{B_m} \ge c_{n,s,p,q,k_0,\lambda,\delta}m^{1/q-1/p}\xrightarrow{m\to \infty} \infty.
\]
Hence, the domain $G_s$ does not support the fractional $(q,p)$-Poincar\'e inequality \eqref{fractionalqp} for any $\tau\in (0,1)$.
\end{proof}
The following theorem shows the
sharpness of Theorem \ref{sharp}.
\begin{theorem}\label{1p_counter}
Let $s>1$, $p\in (1,\infty)$, $\lambda\in [n-1,n)$, and let $\delta,\tau\in (0,1)$.
Suppose that
\[
s<\frac{n+1-\lambda}{1-\delta},\qquad p\le \frac{s(n-1)-\lambda+1}{n-s(1-\delta)-\lambda+1}.
\]
Then, there is an $s$-John domain $G_s$ in $\varmathbb{R}^n$
with the following properties:
$\mathrm{dim}_{\mathcal{M}}(\partial G_s) = \lambda$ and
$G_s$ does not support
the fractional $(1,p)$-Poincar\'e inequality \eqref{fractionalqp}.
\end{theorem}
\begin{proof}
By \cite[Proposition 5.2]{HH-SV} there
is a $1$-John domain $G$ in $\varmathbb{R}^n$
such that
$\mathrm{dim}_{\mathcal{M}}(\partial G)=\lambda$ and
$\limsup_{k\to\infty} 2^{-\lambda k} \cdot \sharp W_k >0$.
By scaling we may also assume that \eqref{w_restr} holds.
Hence, by Proposition \ref{sharp_counter} the $s$-version
$G_s$ has required properties.
\end{proof}
\bibliographystyle{amsalpha}
|
\section{Introduction}
In the standard flare scenario, an acceleration site in the corona (a few Mm to a few tens of Mm above the photosphere),
generates energetic electrons which propagate along magnetic field lines, either towards the lower corona/chromosphere, or into interplanetary space.
As they propagate, they lose energy via Coulomb collisions, and perhaps also via wave-particle interaction.
Energy losses by bremsstrahlung or magnetobremsstrahlung are negligible at our energies of interest ($\approx$1--100 keV).
During close encounters with ambiant ions, electrons emit hard X-rays (HXR) by bremsstrahlung.
There are numerous observations of HXR emission in the solar corona during flares \citep[see e.g.][]{Dennis1985},
but so far none of them have been successfully associated with beams of electrons propagating outwards in the tenuous corona \citep[e.g.][]{Christe2008}.
Unless particle trapping occurs \citep[as is thought to occur in coronal sources, see][and references therein]{Krucker2008},
such observations are indeed difficult to make, given the small column densities electron beams encounter in the corona.
Two sets of independent observations support the existence of outward-going coronal electron beams:
(a) Type III radio bursts which occur simultaneously with HXR emission during the impulsive phase of the flare, and whose frequency decreases with time:
these are interpreted as radiation caused by electron beams which excite plasma emission in the increasingly tenuous coronal plasma as they travel outwards \citep[see e.g.][]{Dulk1985,Bastian1998}.
(b) Interplanetary electrons are detected {\it in situ} at 1 AU \citep[e.g.][]{Lin1985}.
Their onset can often be traced back to a flaring time, when their acceleration is thought to occur \citep{Lin1971,Krucker2007}.
Reconnection theory \citep[ideal MHD, e.g.][and references therein]{Priest2002} generally predicts that magnetic reconnection is symmetrical about the X-point:
Assuming that such reconnection is the mechanism responsible for particle acceleration, then it seems reasonable to expect that
the downward-going beam (that will stop in the chromosphere) and the upward-going beam (that will later escape into interplanetary space) have similar characteristics.
Surprisingly, the number of escaping interplanetary electrons seem to be only about 0.1\%--1\% that of the X-ray producing electrons precipitating in the chromosphere \citep[see e.g.][]{Lin1971,Krucker2007}.
Estimates on the number of electrons required to produce a radio type III burst are difficult to obtain.
\citet{Wentzel1982} used a value of 10$^{33}$ electrons in his discussion of possible theories of Type III bursts.
\citet{Lin1971} have reported 10$^{33}$--10$^{34}$ electrons above 22 keV for interplanetary beams (in situ observations),
and \citet{Kane1972} showed from the upper limit on the flux of thin-target X-rays that less than 10$^{34}$ above 22 keV were required to produce a strong Type III burst at 500 MHz.
There is also evidence \citep[see e.g.][]{Benz1982,Dennis1984} that there exist a secondary acceleration site for Type III-producing electrons,
somehow triggered by the primary energy release: for example, through narrow-band electromagnetic waves from the precipitating flare electrons \citep{Sprangle1983},
or through another, secondary, reconnection process high in the corona \citep{Vrsnak2003}.
The goal of this paper is to numerically estimate the amount (spatial and spectral distribution) of HXRs emitted by electron beams as they propagate,
and determine the ability of various space-borne instruments, namely: RHESSI \citep[Ramaty High Energy Solar Spectroscopic Imager,][]{Lin2002}, GOES, {\it Hinode}/XRT \citep{Golub2007},
and the upcoming FOXSI\footnote{The FOXSI (Focusing Optics X-ray Solar Imager) is a recently accepted rocket flight proposal under the ``Low Cost Access to Space'' (LCAS) NASA program,
which will use grazing incidence mirrors to focus hard X-rays} rocket flight \citep[which will use HXR focusing optics, see e.g.][]{Ramsey2000}, to observe and identify X-ray emission from such electron beams,
at the moment that they exit the presumed acceleration site near the solar surface.
\section{Framework}
The injected (accelerated) beams of electrons are assumed to be power-laws, with a low-energy cutoff $E_1$:
\begin{eqnarray} \label{eq:F0}
F_0(E) & = & \left\{ \begin{array}{ll}
(\delta-1) \frac{F_1}{E_1} \left( \frac{E}{E_1} \right)^{-\delta} & , \,\,\, E > E_1 \\
0 & , \,\,\, E < E_1 \\
\end{array} \right.
\end{eqnarray}
The distribution $F_0(E)$ is expressed in electrons s$^{-1}$ keV$^{-1}$,
$\delta$ is the spectral index, and $F_1$ the total number of injected electrons per second above $E_1$ \citep[same notations as in][]{Brown2002}.
\citet{PSH2002}, have associated the HXR emission from a flare (GOES class C9.6) with a beam of electrons propagating downward, toward the denser chromosphere,
and, assuming a thick-target model \citep{Brown1971}, have found the following characteristics for the injected electron beam:
$\delta$=4, $E_1$=10 keV, and $F_1$=2.7$\times 10^{36}$ electrons/s.
Despite its relatively small X-ray thermal footprint, this flare was particularly hard, and was even a gamma-ray line emitter.
For comparison, the 2002 July 23 flare (GOES class X4.8), had an average electron flux of about 10$^{35}$ electrons/s (electrons above 35 keV) during its $\approx$15-minute long main impulsive phase \citep{Holman2003},
translating to about 10$^{36}$ electrons above 10 keV per second (using an averaged electron spectral index $\delta$ of $\approx$2.5).
In situ and remote observations from \citet{Krucker2007} indicate that the number of electrons in interplanetary beams is $\approx$0.2\% of the number derived from the temporally associated HXR flare beams.
This relationship was established for electrons $>$50 keV.
Assuming it holds for energies down to 10 keV, this means that the interplanetary counterpart of the first flare beam has $F_1\approx$ 10$^{34}$ electrons/s above 10 keV.
We will henceforth call a {\it strong} beam a beam of electrons with $F_{1,strong}$=2.7$\times 10^{36}$ electrons/s, and a {\it weak} beam one with $F_{1,weak}$=1.0$\times 10^{34}$ electrons/s ($F_{1,weak} \approx 0.0037 \times F_{1,strong}$).
A {\it strong} beam is of the type usually associated with flares, whereas a {\it weak} beam is of the type usually associated with Solar Energetic Particles (SEP).
The HXR-producing electron beams in \citet{Krucker2007} had a typical duration of 100 s, and this is the duration that will be used throughout this paper, unless otherwise specified.
For comparison, the 100 s typical duration leads to a total number of {\it weak} beam electrons above 10 keV of $\approx$1$\times$10$^{36}$ electrons, or $\approx$9$\times$10$^{34}$ electrons above 22 keV,
hence about an order of magnitude more than has been reported so far \citep[e.g.][]{Lin1971,Kane1972}, but still below the 4$\times$10$^{36}$ electrons above 10 keV that would have come out of the
2002 July 23 flare, assuming the $\approx$0.2\% relationship holds.
As in \citet{Brown2002} (and using the same assumptions), the electron spectrum changes shape as it propagates, due to Coulomb energy losses, according to:
\begin{equation}
F(E,N)= \frac{E}{\sqrt{E^2 + 2KN}} \,\, F_0 \left( \sqrt{E^2 + 2KN} \right) \\ \label{eq:FEN1}
\end{equation}
\begin{equation}
=\left\{ \begin{array}{ll}
(\delta-1) \frac{F_1}{E_1} E_1^{\delta} \,\, E (E^2 + 2KN)^{-\frac{\delta+1}{2}} & \,\,\, ,E > \zeta \\
0 & \,\,\, ,E < \zeta \\
\end{array} \right.
\end{equation}\\ \label{eq:FEN2}
\noindent where $N$ is the electron column density traversed by the beam of electrons, $K$=2.6$\times 10^{-18}$ cm$^2$ keV$^2$,
and $\zeta$=$\sqrt{\max(0, E_1^2-2KN)}$ is the position of the low-energy cutoff after a column density $N$ has been traversed.
The bremsstrahlung emission per unit column density, along the path of propagation is \citep[from][]{Brown2002}:
\begin{equation} \label{eq:int}
\frac{dI}{dN}(\varepsilon,N) = \frac{1}{4\pi D^2} \int_\varepsilon^\infty F(E,N) \, Q(\varepsilon,E) \, dE
\end{equation}
in photons/s/cm$^{-2}$/keV, where $D$ is 1 AU, $N$ is the column density already traversed by the electron beam,
and $Q(\varepsilon,E)$ is the differential (for emitted photon energy $\varepsilon$) bremsstrahlung cross-section.
Using the Kramers cross-section yields an analytical solution to Eq.~(\ref{eq:int}) \citep[][Appendix~\ref{appendix:dIdN}]{Brown2002}, but does not yield an accurate photon spectrum below the low-energy cutoff.
Hence, numerical evaluations of $\frac{dI}{dN}$ using Eq.~(\ref{eq:int}) and the more proper non-relativistic Bethe-Heitler differential bremsstrahlung cross-section (see appendix \ref{appendix:dIdN}) were used, in order to cover the general case.
Finally, it must be mentioned that we will exclusively use an isotropic bremsstrahlung cross-section.
In reality, bremsstrahlung X-ray emission has anisotropic directivity.
The exact details depend on the spectral shape of the electron energy distribution, the thickness of the target, the pitch angle distribution, the angles between the electron beam trajectory, the magnetic field and the observer, and the energy of the emitted X-rays.
\citet{Elwert1971} (thin-target case) and \citet{Brown1972} (thick-target case, applicable in our case low emitted photon energies: $\varepsilon\lesssim$10 keV),
have both concluded a general limb brightening effect \citep[for more recent work on the topic, see also][]{Massone2004}.
This limb-brightening effect is small at low energies (electron beams at the limb actually produce $\approx$50\% more 10-keV photons than the isotropic bremsstrahlumg cross-section amount),
and increases with photon energy ($\approx$100\% more 50-keV thin-target photons than if assuming an isotropic bremsstrahlung cross-section).
Hence, the effects of bremsstrahlung cross-section anisotropy actually play in our favor, as we are mostly interested in electron beams beyond the solar limb.
\section{Numerical simulations}
The ambiant density is the most important factor contributing to HXR fluxes from beams of electrons in the corona, besides the total number of electrons involved.
It must be of a sufficiently high value.
Type III radio bursts, which are believed to be caused by electron beams through the bump-on-tail plasma instability and Langmuir wave conversion into EM waves,
often start as high as 500 MHz, corresponding to a plasma density of $\sim 3\times10^9$ cm$^{-3}$.
Start ambiant densities for coronal/interplanetary electron beams can hence be inferred to be as high as this
(and even higher, as the electron beam may propagate some distance beyond its injection site before conditions for plasma emission allow the generation of a Type III radio burst).
When ten times the standard Baumbach-Allen \citep{Baumbach1937} coronal density structure is assumed (coronal streamers can be an order of magnitude (or more) denser than the quiet corona \citep{Fainberg1974}),
this density of $\sim 3\times10^9$ cm$^{-3}$ corresponds roughly to an acceleration altitude of 20 Mm (see Fig.~\ref{fig:density}).
Unless otherwise specified, these are the numbers used in our calculations.
\subsection{Spectra, profiles, and RHESSI imaging}
From the injected beam characteristics and the atmospheric density profile, the spatially-integrated photon spectra,
RHESSI count spectra, photon flux spatial profiles,
and FOXSI countrate profiles,
such as in Fig. \ref{fig:sp_prof_Eco10}\footnote{\small Much more at \url{http://sprg.ssl.berkeley.edu/ $\sim$shilaire/work/ebeam\_June2007/wwwoutput/browser2.html}},
can be calculated (a typical flare duration of 100 s was taken to compute the count rates).
In cases where the ``down'' beam is {\it weak} ({\it green curve} in Fig.~\ref{fig:sp_prof_Eco10}) or nonexistant (such as when flare footpoints are occulted by the solar limb),
a {\it strong} ``up'' beam ({\it red curve} in Fig.~\ref{fig:sp_prof_Eco10}) should be easily observed in a RHESSI full-sun spectrum and by FOXSI.
A {\it weak} ``up'' beam ({\it dark purple curve} in Fig.~\ref{fig:sp_prof_Eco10}) is barely noticeable in a RHESSI full-sun spectra, but still well within FOXSI imaging capabilities.
RHESSI simulated imaging (left column pair of Fig.~\ref{fig:rhessiSimImgs}) shows an elongated structure in the case of a {\it strong} ``up'' beam, but very little if the beam is {\it weak}.
The situation improves only marginally in a denser corona (right column pair of Fig.~\ref{fig:rhessiSimImgs}):
the emissivity is indeed increased, but the spatial extend of the emission region is decreased, to a more localized source, as the column density traversed by the beam is thicker.
Going to higher energies is usually of no help, as the fluxes are much smaller (see also Appendix~\ref{appendix:dIdN}).
It is easier to associate an elongated source as coming from a beam of electrons:
a more compact source is more easily associated to an acceleration region or plasmoid with trapped particles.
FOXSI, with its far better dynamic range (50'' away from the main source, its sidelobes are at the 10$^{-3}$ level, as opposed to $\approx$0.1 with RHESSI, S. Christe, PhD Thesis), is much better equipped than RHESSI (typical dynamic range of $\approx$10) to discriminate between these two cases.
The ``up'' beam can clearly be imaged (as an elongated structure) by RHESSI if it is {\it strong}, and if the ``down'' component is {\it weak} or non-existent (occulted).
No such observations have been reported so far, leading to the conclusion that ``up'' beams may very well always be of the {\it weak} kind, i.e. with fluxes $\lesssim$10$^{34}$ (electrons above 10 keV)/s.
It also shows that in partially disk-occulted events, coronal emission produced by a {\it strong} beam in a flare loop is easily observable by RHESSI \citep{Krucker2007a,Krucker2008}.
Because of instrumental sidelobes, the presence of a non-occulted thermal coronal source, such as typically produced by a flare loop,
will completely mask any beams, even if spatially separated, in the case of RHESSI.
Even FOXSI's much-reduced sidelobes will only marginally allow it to observe {\it strong} beams, while {\it weak} beams most assuredly not
(for quick comparison with top plots of Figure~\ref{fig:sp_prof_Eco10}: the thermal flux generated at 10 keV by a typical flare-like 10 MK, 10$^{49}$ cm$^{-3}$ source is about 3$\times$10$^5$ photons s$^{-1}$ cm$^{-2}$ keV$^{-1}$).
\subsection{GOES response}
Table \ref{tab:goesresp} displays the expected response of GOES (GOES 10, both X-ray channels) to the {\it non-thermal} bresstrahlung from our beams of electrons.
Were the ``up'' beams of the {\it strong} kind, GOES should easily observe them.
In case of {\it weak} ``up'' beams, the emission can easily go unnoticed: it lies beneath the digitization level of the instrument (about A0.3 level).
For information, Table \ref{tab:goesTEM} lists the temperature and emission measures that would be derived from the fluxes of Table \ref{tab:goesresp},
using {\it Solarsoft}'s two-filter ratio method.
This method assumes an isothermal plasma, and indicates temperatures of 20--23 MK.
The emission measure results scale well with beam fluxes ({\it weak} being 0.0037 times the {\it strong}) for the Mewe code.
For the Chianti code, there is a software limitation due to the fact that the code is not meant to deal with such low photon fluxes, resulting in inconsistent numbers:
they were hence not displayed for the case of the {\it weak} beam.
For comparison, \citet{Feldmann1996} find in their statistical study that solar flares with such high ($\approx$22 MK) temperatures have an X-ray class of about M3.0,
i.e. emission from both {\it strong} (A4.4) and {\it weak} (A0.02), ``up'' coronal beams, if observed, would stand apart on a (temperature) vs. (GOES X-ray class) plot
(neglecting any local heating by the beams; see Section~\ref{sect:beamheating} for a discussion).
\subsection{Hinode/XRT response}
The new Hinode/XRT \citep{Golub2007} instrument has a whole suite of different filters.
Figure~\ref{fig:XRT_prof} shows the spatial profile from coronal electron beams observed with some of them.
As can be seen in Figure~\ref{fig:BethinSimEbeam}, a {\it weak} ``up'' electron beam is observable with XRT,
provided that a careful choice of image color scale is made.
Optimal conditions are:
\begin{itemize}
\item Long exposures ($>$30 secs)
\item Thin filter (such as Be-thin), for their better spectral response
\item Usage of appropriate image color scales, allowing for certain weak features to be revealed:
even a weak beam-like feature can be distinguishable from the image noise or other dominant features because of its structure in the image (e.g. a straight line,...).
\end{itemize}
\subsection{Effect of beam heating} \label{sect:beamheating}
This short section tries to estimate the effect of heating of the local corona by the non-thermal ``up'' beam:
can the resulting X-ray thermal emission be greater than the X-ray non-thermal emission?
The non-thermal power lost in collisions can in principle heat the local medium:
the amount of non-thermal power dumped along each path element can be calculated.
Given a beam duration $\Delta t$, beam area, and in the absence of thermal losses (from either heat conductivity or thermal radiation), an upper limit to the temperature increase for the ambiant corona along the path of the beam can be determined
(mathematical details in Appendix \ref{appendix:openloopheating}).
As a consistency check on the assumption of no thermal losses, the timescale for heat conductivity losses can be very roughly estimated (see details in Appendix \ref{appendix:closedloopheating}),
using the density scale height of the heated medium.
If this timescale is shorter than the beam duration, then heat losses should be included.
The treatment presented in Appendix \ref{appendix:closedloopheating} is best applied to a closed loop system, but was used as proxy for our case with an open field line,
taking the loop length $L$ to be the density scale height.
In this treatment, the resulting temperature depends weakly on $L$, and our model was deemed sufficiently accurate to provide a rough upper limit of ambiant temperature.
Fig.(\ref{fig:beam_heating}) ({\it bottom}) displays the thermal spectrum expected from our non-thermal beams of electrons,
for both {\it weak} and {\it strong} cases, as well as for different beam areas: (10$^{16}$ cm$^2$ is about 1" radius, and (10$^{18}$ cm$^2$ is for a beam with a radius of about 10" radius.)
In the case of a {\it weak} beam, both conduction and radiative loss timescales were found to be greater than the duration of the beam's injection ($\Delta t$=100 s).
For the {\it strong} beam with a small section ({\it black dotted} line in Fig. \ref{fig:beam_heating}), the heat conduction energy loss timescale (Eq. \ref{eq:tcond}, taking $L$ to be the barometric scale height $H_n$=10$^{10}$ cm) was found to be many orders of magnitude smaller than $\Delta t$,
leading to the use of Eq. (\ref{eq:Teq}) to better estimate the plasma temperature. Table \ref{tab:beamheatingTEM} summarizes the temperatures and emission measures derived.
In the case of a {\it weak} ``up'' beam, it appears that the thermal emission should be negligible in comparison to the beam's non-thermal emission,
at RHESSI energies ($>$3 keV).
At energies below $\sim$3 keV, it is the thermal emission that is expected to dominate.
This result is not very sensitive to the position of the low-energy cutoff (all else, including the total electron flux, being equal).
A {\it strong} ``up'' beam with wide cross-section behaves much the same as the {\it weak} beam cases.
A {\it strong} ``up'' beam with small cross-section, while very hot, is masked by the even stronger non-thermal emission.
To summarize, non-thermal emission are expected to prevail at energies above $\sim$3 keV.
Below $\sim$3 keV, thermal emission is likely to dominate.
Both XRT on {\it Hinode} and SXT on {\it Yohkoh} observe below $\sim$2 keV.
This raises the very interesting possibility that the SXR emission from X-ray jets that are observed by XRT
and previously by SXT, to which Type III radio bursts have been sometimes associated \citep{Shibata1992, Aurass1994, Kundu1995, Raulin1996},
could very well be the direct result of such heating.
X-ray jets on the Sun would then be expected to occur whenever coronal and interplanetary electron beams are observed.
The temperatures (5 MK) and emission measures (10$^{44}$ cm$^{-3}$) obtained by \citet{Kundu1995} for their X-ray coronal jet are qualitatively near those of
Table \ref{tab:beamheatingTEM} for the case of the {\it weak} beam with small cross-section, futher supporting this claim.
A careful search of observations for spatially and temporally correlated HXR, radio Type III, and X-ray jets has been initiated.
\section{Upper limits from observations}
This section includes some observational facts to our discussion so far.
The first part deals with the ``coronal beam associated with X-ray jets'' aspect that was suggested earlier,
followed by a brief discussion on constraints imposed by the oft-observed lack of correlation between
Type III radio bursts and X-ray emission.
\subsection{X-ray jets as coronal beams}
As the previous discussion has suggested that X-ray jets might be associated with coronal electron beams,
we have examined a few polar X-ray jets \citep{Cirtain2007} observed with {\it Hinode} XRT long-exposure images.
Nine have been observed in the period 2007/03/11 21:00 to 2007/03/12 06:00 UT, near the solar limb (but not occulted).
In none of these cases were any X-ray flux enhancements observed with RHESSI, or GOES:
\begin{itemize}
\item
Table~\ref{tab:lcdetection} gives the amount of electron that an {\it up} beam must contain in order to be detectable in spatially-integrated
RHESSI Observing Summary countrates \citep{Schwartz2002}.
The absence of any clear observation puts the upper limit on the total number of electrons in a coronal beam to 0.6$\times$10$^{35}$ electrons above 10 keV
(3--$\sigma$ detection level) over short timescales of a few seconds, and 4.3$\times$10$^{35}$ electrons on timescales of a few minutes.
\item
The GOES low and high channels remained also flat.
For detection by visual inspection of the GOES lightcurves, an increase of $\approx$2$\times$10$^{-9}$ W/m$^2$ in the low (1--8$\AA$) channel
was required over a 3 s time bin (rough estimate).
This corresponds (see Table~\ref{tab:goesresp}) to 6$\times$10$^{35}$ electrons above 10 keV
(1.5$\times$10$^{36}$ electrons in the high (0.5--4$\AA$) channel).
\item One of the X-ray jets (2007/03/12 05:18 UT) lasted about five minutes.
RHESSI imaging over this five minute interval yields no reliable image in the 3--6 or 6--12 keV bands.
As any RHESSI source typically needs at least $\approx$300 counts/detector (empirical value) to be successfully characterized,
this translates into a needed count rate of approximatively 1 counts/s/detector over that time interval.
This puts the needed number of electrons for good imaging to
1.2$\times$10$^{37}$ electrons above 10 keV for the 3--6 keV band, and 3$\times$10$^{36}$ electrons for the 6--12 keV band.
\end{itemize}
Overall, RHESSI lightcurves are more sensitive than GOES lightcurves, and further indicate that observed X-ray jets did not expel more than 0.6$\times$10$^{35}$ electrons above 10 keV
on timescales of a few seconds, and no more than $\approx$5$\times$10$^{35}$ electrons above 10 keV over timescales of a few minutes.
From inspection of Fig.~\ref{fig:sp_prof_Eco10} ({\it bottom left}), about 10$^{33}$ electrons is required for FOXSI imaging.
In cases were the acceleration site is at lower densities than our start density of 3$\times$10$^9$ cm$^{-3}$ (a typical high value),
then these upper limits get proportionally higher. For example, were the start density 3$\times$10$^{8}$cm$^{-3}$,
the upper limit on the number of accelerated electrons is increased tenfold.
\subsection{Type III radio bursts from coronal beams}
The detection thresholds from the previous section can be used again:
For X-ray detection through RHESSI lightcurves, a coronal electron beam would require about 0.6$\times$10$^{35}$ electrons above 10 keV over 4 s,
or about 4$\times$10$^{35}$ electrons above 10 keV over a few minutes.
RHESSI characterization by imaging requires at least 3$\times$10$^{36}$ electrons above 10 keV.
The fact that no clear spatial and temporal correlation of Type III radio burst and non-thermal X-ray emission beyond the limb has ever been established
is already an indicator that electron beams must have typically less than these numbers of electrons.
A systematic search using data from the Nan\c{c}ay Radioheliograph and RHESSI will be initiated shortly.
The best case so far of such an event has been discussed in \citet{Krucker2008}, and mentionned briefly in the conclusion.
\section{Summary and Conclusion}
(1) {\it Strong} escaping (``up'') beams, i.e. with fluxes comparable to the usual chromospheric HXR-producing flare electrons ($\gtrsim$10$^{38}$ electrons above 10 keV),
should easily be detectable and imageable with RHESSI, provided any chromospheric footpoint is occulted.
The absence of such observations supports the scenario established so far, i.e. that escaping electrons are fewer in number than a few tenths of a percent of those hitting the chromosphere.
This in turns hints at asymmetries in the overall standard acceleration scenario.
Possible explanations can range from
(a) the presence of a ``collapsing trap'' mechanism \citep[as described e.g. in][]{Karlicky2004} that enhances the number of accelerated flare electrons,
but not the escaping electrons on open field lines,
(b) the possibility that the main acceleration actually takes place elsewhere than in the high corona, such as in the footpoints, as suggested by \citet{Fletcher2008},
(c) the possibility that escaping electron beams are a secondary energy release phenomenon, triggered by electromagnetic waves from the flare electrons \citep{Sprangle1983},
or (d) the presence of a secondary reconnection process higher up in the corona, where particle densities are much lower, connecting to open field lines \citep[see e.g.][]{Vrsnak2003}.
(2) GOES is not expected to observe anything of note from {\it weak} beams (beams with $\lesssim$10$^{36}$ electrons above 10 keV).
Escaping {\it weak} beams appear to be just below RHESSI's imaging capabilities (even with footpoints occulted),
marginally within Hinode/XRT's imaging capabilities, but well within FOXSI's.
For XRT, thin filters and long exposures are required, as is a careful choice of image dynamic range.
A systematic search of the XRT data, particularly those with long exposure times is currently underway:
For the year 2007, XRT was observing near the solar limb (partial disk images, with image center $>$600" from Sun center) with thin filters and long exposures ($>$30 s) 0.65\% of the time.
NOAA\footnote{ftp://ftp.ngdc.noaa.gov/STP/SOLAR\_DATA/SOLAR\_RADIO/SPECTRAL/SPEC\_NEW.07} reports about 500 different Type III bursts during 2007 (a very quiet year).
Assuming they were produced by beams of electrons that were 30 seconds long, it means that Type III-producing electron beams occur 0.05\% of the time.
The probability for simultaneous occurence of a Type III burst and XRT long-exposure observation with a thin filter is hence p$\approx$3$\times$10$^{-6}$.
Between the start of the Hinode mission, and the end of January 2008, about n=6000 long-exposure pictures with thin filters were taken by XRT.
The chance of there being at least one electron beam caught within that sample can be hence estimated to be 1-(1-p)$^n$ $\approx$ np $\approx$ 2\%,
i.e. we probably haven't observed it.
(3) Coronal emission due to beam heating is not strong enough to mask the non-thermal bremsstrahlung emission of the upgoing beam at energies above 3 keV (i.e. in the enrgy ranges of RHESSI and FOXSI).
With an instrument such as XRT, a rough estimation leads us to expect that the thermal emission might indeed mask the non-thermal emission.
(4) We have raised the possibility that SXR jets might be the result of local heating by the propagating coronal/interplanetary electron beams.
This is consistent with the fact that interplanetary electron beams have been found to be temporally correlated with SXR plasma jets \citep{Wang2006,Pick2006,Nitta2008}.
On the other hand, no non-thermal HXR emission has ever been spatially associated with those SXR jets, probably due to lack of sensitivity.
(5) The absolute minimum amounts of electrons needed for X-ray {\it detection} and for {\it characterization through imaging} are (assuming optimal start densities and minimal backgrounds):
\begin{itemize}
\item $\gtrsim$10$^{35}$ electrons above 10 keV: for detection (and localisation via coarse imaging, to the $\approx arc minute level$) by RHESSI
\item $\gtrsim$6$\times$10$^{35}$ electrons above 10 keV: for detection by GOES
\item $\gtrsim$3$\times$10$^{35}$ electrons above 10 keV: for imaging by Hinode/XRT
\item $\gtrsim$3$\times$10$^{36}$ electrons above 10 keV: for imaging by RHESSI (with sufficient statistics to observe structures to the $\approx$10'' level)
\item $\gtrsim$10$^{33}$ electrons above 10 keV: for imaging by FOXSI (180 cm$^2$ effective area detector, assumes zero background)
\end{itemize}
Appendix~\ref{appendix:requirements} is a list of optimal observations for identification and characterization of escaping coronal electron beams.
An order of magnitude estimate on the expected number of beams with enough electrons to be characterized through RHESSI X-ray imaging can be done as follows:
Assuming that $\approx$10\% of the $\approx$120 electron events per year that WIND observes around solar maximum produce $\approx$10$^{34}$ electrons above 22 keV,
or $\approx$10$^{35}$ electrons above 10 keV, this leads to $\approx$12 events with $\gtrsim$10$^{35}$ electrons above 10 keV per year.
Using the 1.4 power-law negative spectral index found in peak interplanetary electron flux distributions (P.H. Oakley, {\it priv. comm.}),
this leads to an estimate of about 0.5 interplanetary beams with $\gtrsim$10$^{36}$ electrons above 10 keV per solar-maximum year.
With three spacecrafts (WIND, STEREO A \& B), the expectation becomes 1.5 per solar-maximum year.
Periods when occulted flares can be observed from Earth and a spacecraft with in situ intruments magnetically connected to its escaping particles
are around the middle of 2009 with STEREO A and around 2014 with STEREO B.
The best case published so far of an observation of a coronal electron beam, using RHESSI and XRT data, has been discussed in Krucker et al. (2008).
Yet, the XRT coverage was not optimal, no radio imaging was available, and, most importantly,
the number of electrons in the interplanetary beam was at least an order of magnitude below what was inferred from the coronal X-ray emission
(3$\times$10$^{33}$ vs. 10$^{34}$--3$\times$10$^{36}$ electrons above 20 keV).
The author argue that the in-situ measurements, which sample only a very small fraction of the beam's breadth,
might make erroneous assumptions on its spatial distribution, and that in reality many more escaping electrons could be present.
With the latest additions to the fleet of sun-observing spacecraft (STEREO, HINODE, SDO) and the solar activity rising,
it is expected that several such events will be sufficiently observed.
Future spacecraft missions in the inner heliosphere (Solar Orbiter, Sentinels, Solar Probe) will provide regularly such observations at much higher sensitivity.
The scheduled 5-minute FOXSI rocket mission will have the dynamic range and sensitivity required to images faint X-ray emission from outgoing electron beams,
but the chance of observing a radio type III burst during a 5 minute flight is close to zero.
A future space mission with a focusing optics telescope dedicated to solar observations, however, would revolutionize our understanding of electron acceleration in solar flares.
The Nuclear Spectroscopic Telescope Array (NuSTAR) Small Explorer satellite \citep{Harrison2005}, to be launched in 2011, uses HXR focusing optics for astrophysical observations with an effective area of 1000 cm$^2$.
With a solar mission of similar size as NuSTAR, HXR emission from escaping electron beams will be generally detected with excellent statistics allowing us to spectrally image the electron acceleration region and trace electron beams from their acceleration region down to the chromosphere as well as into interplanetary space.
\acknowledgments
This work was supported by NASA Heliospheric GI awards NNX07AH74G and NNX07AH76G, and by Swiss National Foundation (SNSF) grant PBEZ2-108928.
We would like to thank the anonymous referee for his constructive comments.
{\it Facilities:} \facility{RHESSI}, \facility{Hinode (XRT)}, \facility{GOES}, \facility{FOXSI}.
|
\section{Introduction}
Stable massive particles (SMPs) offer spectacular detector signatures entirely without physics backgrounds.
This paper summarizes the results of two searches for SMPs~\cite{Aad:2011yf,Aad:2011hz} performed with the ATLAS experiment
at the LHC using $34$-$37$~pb$^{-1}$ from 2010.
\section{Outline of the searches}
Lepton-like SMPs, such as long-lived sleptons, are expected to lose energy through EM processes and give rise to a detector signature similar to that of slow-moving muons. Hadron-like SMPs, such as $R$-hadrons, could exchange electric charge when penetrating the detector material and therefore possibly be dominantly neutral in the inner tracker or the muon spectrometer~\cite{deBoer:2007ii,Farrar:2010ps,Mackeprang:2009ad}.
Therefore, two strategies were pursued.
The first search~\cite{Aad:2011yf} targets hadron-like SMPs uses only the inner detector (ID) and calorimetry in ATLAS, without relying on the muon spectrometer (MS). In events triggered by a $E_{\mathrm{T}}^{\mathrm{miss}} > 40$~GeV signature, slow-particle candidates are sought as tracks with $p_{\mathrm{T}} > 50$~GeV in $|\eta| < 1.7$. The speed of the candidate is then estimated in two independent ways. Through the Bethe-Bloch relation $\beta\gamma$ is extracted from the ionization energy loss measurement in the Pixel detector, and $\beta$ is estimated from time-of-flight measurements in the Tile calorimeter. By combining the track momentum and speed measurements, two mass estimates are obtained through \mbox{$m = p/\beta\gamma$}. The signal regions are defined in the resulting two-dimensional mass plane, and a data-driven method is used to estimate the yields from background processes.
The second search~\cite{Aad:2011hz} relies on a signal in the MS.
A muon trigger was used to collect the experimental data, and two selections were defined to target lepton-like and hadron-like SMPs separately. For the lepton-like SMPs, two candidates with $p_{\mathrm{T}} > 40$~GeV are required in the events, each with a combined track in both the ID and MS. For the hadron-like SMP selection only one candidate with $p_{\mathrm{T}} > 60$~GeV is required per event, and in the absence of combined tracks standalone MS tracks are also used, in order to be sensitive to $R$-hadrons that are neutral in the ID and charged in the MS. One single $\beta$ measurement is then formed using information from both the monitored drift tube precision chambers and the fast resistive plate trigger chambers in the muon spectrometer, as well as the Tile calorimeter where available. Combined with the track momentum measurement, a mass estimate can be calculated for each candidate. A data-driven background estimate was also used for this search.
\section{Results}
In both searches the observations agree well with the low predicted yields due to instrumental effects in a background-only hypothesis, and 95\% C.L. cross section limits are calculated for the production cross section of each signal scenario. Figure~\ref{fig:limits} illustrates the extracted cross section limits along with theoretical predictions for production of $\tilde{g}$, $\tilde{t}$ and $\tilde{b}$ (left), and a model of gauge-mediated supersymmetry breaking (GMSB) featuring a stable $\tilde{\ell}$ (right). The intersections give the resulting mass constraints: $m_{\tilde{g}} > 586$~GeV, $m_{\tilde{t}} > 309$~GeV, $m_{\tilde{b}} > 284$~GeV and $m_{\tilde{\tau}} > 136$~GeV.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.56\textwidth]{combined_limit}
\includegraphics[width=0.41\textwidth]{MS_GMSB_limit}
\caption{Measured 95\% C.L. upper cross section limits for $R$-hadrons (left) and long-lived sleptons (right), along with predictions based on theoretical calculations.}
\label{fig:limits}
\end{center}
\end{figure}
|
\section{The PREX experiment, neutron densities, and neutron radii}
\label{Sec1}
Nuclear charge densities have been accurately measured with electron scattering and have become our picture of the atomic nucleus, see for example ref. \cite{chargeden}. These measurements have had an enormous impact.
In contrast, our knowledge of neutron densities comes primarily from hadron scattering experiments involving for example pions \cite{pions}, protons \cite{protons1,protons2,protons3}, or antiprotons \cite{antiprotons1,antiprotons2}. However, the interpretation of hadron scattering experiments is model dependent because of uncertainties in the strong interactions.
Parity violating electron scattering provides a model independent probe of neutron densities that is free from most strong interaction uncertainties. This is because the weak charge of a neutron is much larger than that of a proton \cite{dds}. Therefore the $Z^0$ boson couples primarily to neutrons. In Born approximation, the parity violating asymmetry $A_{pv}$, the fractional difference in cross sections for positive and negative helicity electrons, is proportional to the weak form factor. This is very close to the Fourier transform of the neutron density. Therefore the neutron density can be extracted from an electro-weak measurement \cite{dds}.
Many details of a practical parity violating experiment to measure neutron densities have been discussed in a long paper \cite{bigprex}.
The doubly magic nucleus $^{208}$Pb has 44 more neutrons than protons, and some of these extra neutrons are expected to be found in the surface where they form a neutron rich skin.
The thickness of this skin is sensitive to nuclear dynamics and provides fundamental nuclear structure information. There may be a useful analogy with cold atoms in laboratory traps were similar ``spin skins'' have been observed for partially polarized systems \cite{coldatomsMIT, coldatomsRICE}. Note that there is an attractive interaction between two atoms of unlike spins while for a nucleus, the interaction between two nucleons of unlike isospins is also more attractive than the interaction between two nucleons of like isospins.
The neutron radius of $^{208}$Pb, $R_n$, has important implications for astrophysics. There is a strong correlation between $R_n$ and the pressure of neutron matter $P$ at densities near 0.1 fm$^{-3}$ (about 2/3 of nuclear density) \cite{alexbrown}. A larger $P$ will push neutrons out against surface tension and increase $R_n$. Therefore measuring $R_n$ constrains the equation of state (EOS) --- pressure as a function of density --- of neutron matter.
Recently Hebeler et al. \cite{hebeler} used chiral perturbation theory to calculate the EOS of neutron matter including important contributions from very interesting three neutron forces.
From their EOS, they predict $R_n-R_p= 0.17 \pm 0.03$ fm. Here $R_p$ is the known proton radius of $^{208}$Pb. Monte Carlo calculations by Carlson et al. also find sensitivity to three neutron forces \cite{MC3n}. Therefore, measuring $R_n$ provides an important check of fundamental neutron matter calculations, and constrains three neutron forces.
The correlation between $R_n$ and the radius of a neutron star $r_{NS}$ is also very interesting \cite{rNSvsRn}. In general, a larger $R_n$ implies a stiffer EOS, with a larger pressure, that will also suggest $r_{NS}$ is larger. Note that this correlation is between objects that differ in size by 18 orders of magnitude from $R_n\approx 5.5$ fm to $r_{NS}\approx 10$ km.
The EOS of neutron matter is closely related to the symmetry energy $S$.
This describes how the energy of nuclear matter rises as one goes away from equal numbers of neutrons and protons. There is a strong correlation between $R_n$ and the density dependence of the symmetry energy $dS/dn$, with $n$ the baryon density. The symmetry energy can be probed in heavy ion collisions \cite{isospindif}. For example, $dS/dn$ has been extracted from isospin diffusion data \cite{isospindif2} using a transport model.
The symmetry energy $S$ helps determine the composition of a neutron star. A large $S$, at high density, implies a large proton fraction $Y_p$ that will allow the direct URCA process of rapid neutrino cooling. If $R_n-R_p$ is large, it is likely that massive neutron stars will cool quickly by direct URCA \cite{URCA}. In addition, the transition density from solid neutron star crust to the liquid interior is strongly correlated with $R_n-R_p$ \cite{cjhjp_prl}.
Recently, Reinhard and Nazarewicz claim that $R_n-R_p$ is strongly correlated with the dipole polarizability $\alpha_D$ of $^{208}$Pb \cite{Reinhard} and Tamii et al. have accurately determined $\alpha_D$ from very small angle proton scattering \cite{Tamii}. However, further work suggests the correlation of $\alpha_D$ with $R_n-R_p$ is model dependent \cite{JP2011}.
Finally, atomic parity violation (APV) is sensitive to $R_n$ \cite{pollockAPV},\cite{brownAPV},\cite{bigprex}. Parity violation involves the overlap of atomic electrons with the weak charge of the nucleus, and this is primarily carried by the neutrons. Furthermore, because of relativistic effects the electronic wave function can vary rapidly over the nucleus. Therefore, the APV signal depends on where the neutrons are and on $R_n$. A future low energy test of the standard model may involve the combination of a precise APV experiment along with PV electron scattering to constrain $R_n$. Alternatively, measuring APV for a range of isotopes can provide information on neutron densities \cite{berkeleyAPV}.
The Lead Radius Experiment (PREX) measures the parity violating asymmetry $A_{pv}$ for 1.05 GeV electrons scattering from $^{208}$Pb at five degrees \cite{prex}. This measurement initially ran at Jefferson Laboratory in the spring of 2010. PREX demonstrates this new technique to measure neutron densities and achieved the desired small systematic errors. Therefore the measurement can be improved by accumulating more statistics. The collaboration purposes to design appropriate engineering modifications to the beamline to mitigate radiation problems and has been granted additional beam time to improve the statistics and achieve the original goal of a 1\% ($\pm 0.05$ fm) constraint on the neutron radius of $^{208}$Pb.
In addition to PREX, many other parity violating measurements of neutron densities are possible, see for example \cite{PREXII}. Neutron radii can be measured in many stable nuclei, as long as the first excited state is not too low in energy (so that elastically scattered electrons can be separated from inelastic excitations). In general, $R_n$ is easier to measure in lighter neutron rich nuclei. This is because $R_n$ is smaller and so can be measured at higher momentum transfers where $A_{pv}$ is larger.
Measuring $R_n$ in $^{48}$Ca is particularly attractive. First, $^{48}$Ca has a higher experimental figure of merit than $^{208}$Pb. Therefore a $^{48}$Ca measurement may take less beam time than for $^{208}$Pb. Not only does $^{48}$Ca have a large neutron excess, it is also relatively light. With only 48 nucleons, microscopic coupled cluster calculations \cite{coupledcluster}, or no core shell model calculations \cite{NCSM}, may be feasible for $^{48}$Ca that are presently not feasible for $^{208}$Pb. Note that these microscopic calculations may have important contributions from three nucleon forces. This will allow one to make microscopic predictions for the neutron density and relate a measured $R_n$ to three nucleon forces and in particular to very interesting three neutron forces.
We end this section with a discussion of perhaps the ultimate neutron density measurement. Measuring $A_{pv}$ for a range of momentum transfers $q$ allows one to directly determine the neutron density $\rho_n(r)$ as a function of $r$. For $^{208}$Pb this may be extraordinarily difficult. However, for $^{48}$Ca this may actually be feasible, although somewhat difficult and time consuming. For example, one might be able to determine about 6 Fourier Bessel coefficients in an expansion of $\rho_n(r)$. Note that this may require long runs and careful control of backgrounds and could be helped by using a large acceptance spectrometer. Nevertheless, one should not underestimate the utility of having model independent determinations of both the neutron and proton densities as a function of $r$. This will literally provide our picture of the neutrons and protons inside an atomic nucleus.
\section{Radiative corrections to PREX and QWEAK}
\label{Sec2}
Radiative corrections involving additional photons are important for a variety of precision electromagnetic and electroweak experiments. Coulomb distortion and dispersion corrections are illustrated in Fig. \ref{Fig2}. Coulomb distortions, Fig. \ref{Fig2} (b), involve intermediate states where the nucleus remains in its ground state, while dispersion corrections, Fig. \ref{Fig2} (c), involve intermediate states where the nucleus is excited.
\begin{figure}[h]
\center\includegraphics[width=4in]{horowitz_fig1.pdf}
\caption{\label{Fig2} Diagrams describing the electroweak scattering of an electron (top line) with a nucleus (bottom thick line), see text.}
\end{figure}
The individual protons contribute coherently to Coulomb distortions, and as a result Fig. \ref{Fig2} (b) is order $Z\alpha$ compared to the Born term in Fig. \ref{Fig2} (a). Here $Z$ is the atomic number and $\alpha$ is the fine structure constant. Coulomb distortions are clearly important for PREX because of the large $Z=82$ of Pb. Coulomb distortions can be included to all orders in $Z\alpha$ by numerically solving the Dirac Equation for an electron moving in the Coulomb potential of order 25 MeV and a weak axial vector potential of order 10 eV \cite{coulombdistortions}. Coulomb distortions reduce $A_{pv}$ by about 30\% and are the largest known correction to the Born approximation. However the uncertainty in this correction is very small because the charge density of $^{208}$Pb is accurately known. The observed $A_{pv}$ is consistent with Coulomb distortion calculations and inconsistent with Born approximation predictions for any conceivable neutron density. Therefore, PREX provides the first observation of Coulomb distortion effects in parity violating electron scattering.
In general the individual protons contribute incoherently to dispersion corrections. Therefore Fig. \ref{Fig2} (c) is of order $\alpha$ rather than $Z\alpha$ and not obviously large for PREX. There is some information on dispersion contributions to elastic electron scattering, see for example \cite{dispersion}. However, dispersion corrections may be particularly important for parity violating electron scattering from the proton because the Born contribution in Fig. \ref{Fig2} (a) is small, of order $Q_W^p \approx 0.07$, where $Q_W^p$ is the small weak charge of the proton. As a result dispersion corrections in Fig. \ref{Fig2} (c) can make a relative contribution of order $\alpha/Q_W^p\approx 10\%$. Indeed Gorchtein first calculated a large dispersion correction to the weak charge of the proton that should be important for the interpretation of the QWEAK experiment \cite{G2009}. Recently Gorchtein, Horowitz, and Ramsey-Musolf calculated this correction in more detail and estimated the remaining theoretical uncertainty \cite{G2011}. About half of the correction involves nucleon resonance excited states while the remainder is from higher energy ``background'' excitations. An important remaining uncertainty is the isospin decomposition of this background. This is necessary to predict the electroweak response from purely electromagnetic data \cite{G2011}. The dispersion correction $\Delta R$ predicted by \cite{G2011} to the weak charge of the proton is,
\begin{equation}
\Delta R =[5.39\pm0.27\,(mod.\,avg.)\pm1.88\,(backgr.)^{+0.58}_{-0.49}\,(res.)\pm0.07\,(t-dep.)]\,10^{-3}.
\end{equation}
This is at the kinematics of the QWEAK experiment, $E=1.165$ GeV, and a squared momentum transfer of $t=-0.03$ GeV$^2$. The first error involves the model used to fit electromagnetic data, while the second (largest) error involves uncertainties in high energy ``background'' contributions including their isospin dependence. Finally smaller errors are quoted for the resonance contributions $(res.)$ and for the extrapolation from $t=0$, where the calculation was performed, to $t=-0.03$ GeV$^2$. This correction $\Delta R$ is to be compared to the total weak charge of the proton $Q_W^p=0.0767\pm 0.0008\pm 0.0020$, for a relative contribution of $\Delta R/Q_W^p=7.6\pm 2.8 \%$, \cite{G2011}. The correction is large compared to the desired 2\% statistical error for QWEAK. Therefore it must be accurately included in the interpretation of experimental results. The theoretical uncertainty can be reduced by measuring ``deep'' inelastic parity violating electron scattering at high excitation energies and moderate momentum transfers to constrain the isospin of background contributions. Some other calculations of $\Delta R$ include \cite{A2011,B2011,R2011}.
At lower energies $\Delta R$ is both smaller and has a much smaller uncertainty. At $E=0.180$ GeV, $t=0$ we calculate \cite{G2011},
$\Delta R =[1.32\pm0.05\,(mod.\,avg.)\pm0.27\,(backgr.)^{+0.11}_{-0.08}\,(res.)]\,10^{-3}$.
Therefore a low energy QWEAK like experiment, see for example \cite{MAINZ}, should be very cleanly interpretable in terms of the weak charge of the proton.
\acknowledgments
This work was done in collaboration with Mikhail Gorchtein and Shufang Ban and was supported in part by DOE grant DE-FG02-87ER40365.
|
\section{Introduction}
\label{}
A quarter century after the discovery of high-temperature superconductivity \cite{bedn86}, the nature of hole-doped copper-oxide compounds remains controversial. The theoretical machinery that has been developed to describe conventional superconductors is built on top of Fermi liquid theory, but there is considerable experimental evidence suggesting that the normal state properties of cuprates are inconsistent with Fermi liquid predictions over much of the interesting part of the phase diagram. The usual starting point of band theory, in which minimizing the kinetic energy of the conduction electrons plays a dominant role, is inadequate. Electron-electron interactions play a crucial role; however, they are not so strong that one can apply a perturbation theory about the strong-coupling limit. Interactions and kinetic energy are roughly comparable at the doping levels where superconductivity occurs, and this intermediate-coupling regime poses a particular challenge to theory.
Experiments indicate that the doped cuprates have a mixture of characters. To see this, we start with an undoped parent compound such as La$_2$CuO$_4$. This material is an insulator with a charge-transfer gap of $\sim2$~eV. Antiferromagnetism develops within the CuO$_2$ planes as a consequence of strong onsite Coulomb repulsion between electrons in the same Cu $3d_{x^2-y^2}$ orbital \cite{kast98}. The effective magnetic interaction is well characterized by the superexchange mechanism \cite{ande87}, and the magnetic excitation spectrum is described quite well by spin-wave theory with nearest-neighbor superexchange energy $J\sim 140$~meV \cite{head10}.
Things start to change as soon one begins to dope holes into the planes. In La$_{2-x}$Sr$_x$CuO$_4$, the long-range antiferromagnetic (AF) order is destroyed by $x=0.02$, to be replaced by a spin-glass phase with incommensurate magnetic order that develops below 10~K; furthermore, there is evidence for phase separation between the AF and incommensurate phases for $x<0.02$ \cite{mats02}. The rapid destruction of AF order as mobile charge carriers are introduced indicates a competition between the tendency of the holes to delocalize, in order to reduce their kinetic energy, and the onsite Coulomb interactions that drive the AF correlations. Part of this effect can be understood if we think about a single-band model (Cu-sites only) and consider an individual hole moving in an AF background \cite{trug88,lau11}. An electron near the Fermi energy moving in the so-called ``nodal'' direction can hop on the same AF sublattice; no spin flips are involved, so there is no conflict with the AF correlations. In contrast, an electron hopping in the antinodal direction, along Cu-O bonds, can only do so by flipping spins and disrupting the AF correlations. The impact of the doped holes on the AF background becomes even more obvious when one takes account of the fact that doped holes have strong O $2p$ character, as a hole localized between on an O between a pair of Cu ions will induce a ferromagnetic alignment of those Cu spins \cite{emer88,ahar88}. In a different direction, there have been proposals that doping should simply cause the spin correlations to change to a spiral form \cite{lusc07}; however, in that case, one would not expect to see such a dramatic loss of order at $x=0.02$.
Neutron scattering experiments show that dynamical AF correlations survive in the doped cuprates throughout most of the superconducting range \cite{fuji11}, but how can a symptom of the correlated insulator state coexist with itinerant charge carriers?
\section{Stripe order}
One way for locally-AF spin correlations to coexist with mobile holes is through the formation of charge and spin stripes. These stripe patterns are easiest to analyze when they are statically ordered. Theoretical motivations for stripes have been reviewed in \cite{kive03,vojt09,zaan01}. Experimentally, charge and spin stripe order is actually quite common in layered, transition-metal-oxide compounds such as La$_{2-x}$Sr$_x$NiO$_4$\ \cite{yosh00} and La$_{2-x}$Sr$_x$CoO$_4$ \cite{saki08,cwik09}. Of course, these latter systems tend to be insulating when stripe ordered. A key difference in the cuprates is that magnetic Cu ions have a single unpaired $3d$ electron, with spin $S=1/2$, which is essential to the mobility of the doped holes \cite{tran98c}.
The first indication of an anomaly possibly associated with stripes was the discovery of a sharp depression in the superconducting transition temperature, $T_c$, centered on the doping level $x=1/8$ in La$_{2-x}$Ba$_x$CuO$_4$\ (LBCO) \cite{mood88}, an effect not observed (or, at least, not as strongly) in La$_{2-x}$Sr$_x$CuO$_4$\ (LSCO). An x-ray diffraction study by Axe {\it et al.} \cite{axe89} demonstrated that La$_{2-x}$Ba$_x$CuO$_4$\ exhibits a subtle structural phase transition at low temperatures that causes orthogonal Cu-O bonds within each plane to become inequivalent. (The orientation of the anisotropy rotates $90^\circ$ from layer to layer.) This structural anisotropy appears to be important to the development of static stripe order, which was first detected by neutron diffraction in the isostructural compound La$_{1.48}$Nd$_{0.4}$Sr$_{0.12}$CuO$_4$ \cite{tran95a}. It proved challenging to grow La$_{2-x}$Ba$_x$CuO$_4$\ crystals at the same hole concentration, but eventually Fujita and collaborators \cite{fuji04} were successful, allowing confirmation of stripe order in La$_{1.875}$Ba$_{0.125}$CuO$_4$.
The phase diagram for charge and spin stripe order in La$_{2-x}$Ba$_x$CuO$_4$\ has now been established as a function of doping through a combination of neutron and x-ray diffraction, as well as magnetic susceptibility measurements \cite{huck11}. While the amplitude of the stripe order is greatest at $x=1/8$, weak charge stripe order is still detectable at $x=0.095$ and 0.155. Measurements by other groups \cite{duns08,adac01} are consistent with the phase diagram after one calibrates the relative compositions through the doping dependence of structural phase transition temperatures \cite{huck11}. The maximum amplitude of the stripe order is correlated with a minimum of $T_c$ for bulk superconductivity; however, there is evidence of superconducting correlations at much higher temperatures \cite{li07,tran08}, as will be discussed.
An important question concerns the strength of the stripe order. Does it involve substantial local magnetic moments or a weak spin-density modulation? How big is the charge modulation? One measure of the moment size is given by muon spin rotation measurements, which probe the local hyperfine field at the muon site. Analysis of such measurements indicates a maximum ordered moment at low temperature of $\sim0.3$~$\mu_{\rm B}$, about 60\%\ of the value in AF La$_2$CuO$_4$ \cite{nach98}; this is a substantial value, considering the importance of quantum fluctuations. There have also been measurements of the anisotropic bulk susceptibility on a single crystal \cite{huck08}. Below the spin-stripe-ordering transition, a large temperature-dependent anisotropy of the susceptibility is found, consistent with what one would expect to see in an AF insulator. With the field aligned along a Cu-O bond direction, a spin-flop transition is observed at $\sim6$~T \cite{huck08}, again consistent with behavior typically found in systems where a local-moment description is appropriate.
Regarding the charge, neutron and hard-x-ray diffraction are sensitive just to atomic displacements, which provide only an indirect measure of charge modulation. Abbamonte {\it et al.}\cite{abba05} used soft x-ray scattering to demonstrate that the charge-order diffraction peak is resonant at the energy of the O $2p$ pre-edge peak in the O $K$-edge absorption spectrum. This directly demonstrates that the occupancy of the O $2p$ states is spatially modulated with the period of the charge stripes. Quantitative analysis indicated a substantial modulation amplitude \cite{abba05}, consistent with a mean-field calculation by Lorenzana and Seibold \cite{lore02}. Such resonant scattering measurements have since been used to determine the charge-ordering phase diagram for La$_{1.8-x}$Eu$_{0.2}$Sr$_x$CuO$_4$ \cite{fink11}, and to confirm the charge-stripe order in La$_{1.48}$Nd$_{0.4}$Sr$_{0.12}$CuO$_4$ \cite{wilk11}.
Weak incommensurate spin order has also been observed in rather underdoped YBa$_2$Cu$_3$O$_{6+x}$\ \cite{haug09}, and this order is enhanced by an applied magnetic-field. At slightly larger hole doping, similar order can be induced by substituting Zn for 2\%\ of the Cu atoms \cite{such10}. While the
observed incommensurability at a given hole concentration is systematically smaller than that observed in La$_{2-x}$Ba$_x$CuO$_4$\ \cite{haug10}, the trend with doping and the orientation of the modulation wave vector is quite similar. Very recently, Wu {\it et al.} \cite{wu11} have reported evidence from nuclear magnetic resonance measurements for charge stripe order at a hole concentration near 1/8 induced by magnetic fields greater than 25~T applied perpendicular to the planes.
\section{Stripe phase is 2D}
While stripe order involves a unidirectional modulation, the electronic character of the stripe-ordered layer is two-dimensional (2D). Model calculations \cite{gran08} indicate that the Fermi surface corresponding to dispersion along the charge stripes consists of flat sections in the antinodal $(\pi,0)$ and $(0,\pi)$ regions of reciprocal space, while the states along the Fermi arc, extending about $(\pi/2,\pi/2)$, are much more homogeneous in terms of distribution in real space. It has been pointed out that the Fermi arc states have dominant oxygen character, while the antinodal states have more copper character \cite{niks11}.
Experimental angle-resolved photoemission spectroscopic (ARPES) studies on stripe-ordered\linebreak
La$_{1.875}$Ba$_{0.125}$CuO$_4$\ display behavior very similar to that seen in other cuprate superconductors \cite{vall06,he09}. The spectral function measured vs.\ energy shows sharp peaks along the Fermi arc, but much broader features in the antinodal region. A $d$-wave like gap is found along the Fermi arc for $T\lesssim 40$~K, while a somewhat larger gap is present in the antinodal states. The Fermi arc and antinodal gap have also been observed in stripe-ordered La$_{1.48}$Nd$_{0.4}$Sr$_{0.12}$CuO$_4$ \cite{chan08b}. Despite the coexisting magnetic order, there are no obvious features in the ARPES spectra of either of these systems that would indicate the presence of stripe order.
Transport properties are also consistent with the stripe-ordered state remaining effectively metallic. The in-plane resistivity is dominated by states near the Fermi arc, and in La$_{1.875}$Ba$_{0.125}$CuO$_4$\ it retains a metallic temperature derivative below the charge-ordering temperature $T_{\rm co}$\ \cite{li07}. It should be noted that Adachi {\it et al.} \cite{adac11} have seen an upturn in the in-plane resistivity on cooling below $T_{\rm co}$; however, the measurements are challenging due to the extreme anisotropy of the electronic properties \cite{wen11}. Optical conductivity measurements with in-plane polarization show a narrowing of the Drude peak just below $T_{\rm co}$, consistent with metallic behavior \cite{home06,home11}. Low-temperature metallic behavior is not universal for all stripe-ordered cuprates. For example, significant upturns in the in-plane resistivity have been observed in La$_{1.6-x}$Nd$_{0.4}$Sr$_x$CuO$_4$\ \cite{ichi00,daou09b} and there is an absorption peak in the low-frequency optical conductivity for La$_{1.275}$Nd$_{0.6}$Sr$_{0.125}$CuO$_4$ \cite{dumm02}.
\section{Diffraction vs.\ scanning tunneling spectroscopy}
Real-space modulations of tunneling conductance have been imaged on cleaved samples of Bi$_2$Sr$_2$CaCu$_2$O$_{8+\delta}$\ (Bi2212) and
\linebreak
Bi$_{2-y}$Pb$_y$Sr$_{2-z}$La$_z$CuO$_{6+x}$ (Bi2201) \cite{howa03b,kohs07,lawl10,park10,wise08}. The modulations have a period of $\sim4$ lattice spacings, which is similar to the charge stripe period in the $n_h=1/8$ phase; however, the STS modulation wave vector decreases with doping \cite{wise08}, whereas the stripe wave vector increases with doping, before saturating above $n_h=1/8$ \cite{birg06}. Now, these results have been measured in different systems, so they are not in direct conflict; however, unpublished data on Bi2201 from Fujita, Enoki, and Yamada indicate that the doping dependence of the spin-stripe wave vector is actually quite similar to that in LSCO. This leads to a challenge in how to reconcile the results of these distinct techniques.
It has been proposed, based on ARPES work, that the modulation seen by STS corresponds to $2k_{\rm F}$ scattering determined by the parallel Fermi surface sheets in the antinodal regions \cite{shen05}. This matches fairly well in magnitude and doping dependence. On the other hand, in a recent ARPES study \cite{gweo11}, it has been argued that there is a small but systematic discrepancy between these quantities. It should be noted that the ``stripe''-like modulation signal in STS is strongest at a substantial bias voltage, on the order of the pseudogap energy, and the relevant wave vector at that energy might be shifted from that at the Fermi energy (zero bias voltage). Furthermore, the pseudogap energy is measured from these antinodal states, so it seems likely that they are connected.
If STS is detecting a $2k_{\rm F}$-like modulation, it is not incompatible with stripes. As discussed above, charge stripes should have Fermi surface segments in the antinodal regime \cite{gran08}; however, the associated modulation would be orthogonal to the charge-stripe modulation. Parker {\it et al.} \cite{park10} have shown in Bi2212 that the STS modulation strength and temperature onset maxima occur close to $n_h=1/8$, supporting a connection to charge stripes.
\section{Stripe dynamics}
Spin fluctuations disperse from the incommensurate magnetic superlattice peaks in the stripe ordered phase \cite{fuji04,duns08,tran99a}. In La$_{1.875}$Ba$_{0.125}$CuO$_4$, the spectrum has been measured up to $\sim200$~meV \cite{tran04}, and it is found to share the same ``hour-glass" dispersion as other cuprate superconductors \cite{fuji11}. The energy $E_{\rm cross}$ of the crossing point of the spectrum varies linearly with doping on the underdoped side, decreasing towards zero as the doping is reduced. Thus, the upwardly dispersing excitations evolve into the spin waves of the AF phase. The spectrum below $E_{\rm cross}$ is strongly modified from the AF behavior because of the presence of the doped holes. Many researchers have attempted to explain the downwardly dispersing spectrum in terms of particle-hole excitations of a homogeneous system \cite{esch06}; however, such a mechanism is challenged when it comes to explaining a number of the features observed in LBCO.
In La$_{1.875}$Ba$_{0.125}$CuO$_4$, the low-energy incommensurate spin fluctuations survive above $T_{\rm co}$, indicating the presence of fluctuating stripes \cite{xu07}. These low-energy excitations are very similar to those in LSCO throughout the underdoped regime, suggesting that fluctuating stripes are a common feature.
At modest energies, the magnetic spectral weight in the doped cuprates is comparable to that of AF spin waves, but Stock {\it et al.} \cite{stoc10} have pointed out that the spectral weight is suppressed above the pseudogap energy. This observation suggests that particle-hole excitations compete with spin fluctuations, rather than reinforcing them. Stripe-like segregation of spins and holes provides a mechanism to maintain AF correlations with local-moment character. There must be a balance between the AF energy and the kinetic energy of the holes, so it is reasonable to expect that the stripe-like correlations will eventually disappear at sufficient hole density.
\section{Stripes and superconductivity}
While stripe order competes with bulk superconducting order, evidence has been found for strong two-dimensional superconducting (SC) correlations that begin at a temperature comparable to the typical bulk $T_c$ \cite{li07,tran08}. It is quite unusual to observe 2D SC in a bulk crystal, as interlayer Josephson coupling inevitably leads to 3D SC order. To explain the decoupling, a pair density wave (PDW) SC state has been proposed \cite{hime02,berg07}. In the PDW state, the pair wave function is locally $d$-wave-like, but it is modulated by an envelope function that oscillates sinusoidally, with extrema aligned with the charge stripes, and nodes centered in the spin stripes.
Berg {\it et al.} \cite{berg09a} considered mechanisms that would induce the antiphase coupling of neighboring SC stripes. A couple of recent calculations based on the 2D $t$-$J$ model have identified conditions where the PDW phase appears to be the ground state. Loder {\it et al.} \cite{lode11} have analyzed mean field models valid for either small or large $J/t$, while Corboz {\it et al.} \cite{corb11} made use of a new type of variational Monte Carlo scheme for an intermediate value of $J/t$. In the latter calculation, there is no significant difference in energy between in-phase and antiphase coupling between superconducting stripes.
Although stripe order can be bad for SC phase coherence, it appears to be compatible with strong pairing. Kivelson and Fradkin \cite{kive07} have argued that stripe-like inhomogeneity can enhance pairing and superconductivity. Empirically, a close relationship between stripes and superconductivity is suggested by the fact that stripe ordering temperatures are always close to $T_c$ values. Dynamic fluctuations of the inhomogeneity may be essential for optimizing the phase order at high temperature.
\section{Magnetic-field-induced stripe order}
It has been known for some time that spin stripe order can be induced by a $c$-axis magnetic field \cite{lake02}. Recent measurements of LBCO with $x=0.095$, which has only very weak stripe order in zero field \cite{huck11}, have shown for the first time that charge stripe order can also be enhanced by a field \cite{wen11}. The presence of the stripe order appears to weaken the interplanar Josephson coupling, but does not reduce it to zero. Transport measurements indicate that the $c$-axis field can can cause the interlayer resistivity to become finite while the in-plane resistivity remains zero \cite{wen11}. (Such behavior violates conventional theoretical expectations, as one expects for a layered superconductor in a magnetic field that there is superconducting phase order in all three dimensions or in none.) The condition of zero resistivity for currents parallel to the planes is maintained to much higher fields and temperatures than found for comparably doped LSCO \cite{sasa00,gila05}. Thus, it appears that the special crystal structure of LBCO is capable of pinning the magnetic vortices via the induced stripe order, limiting dissipative flux flow, though not preventing slips in the phase of the superconducting order parameter when currents flow along the $c$ axis.
\section{Acknowledgements}
I am grateful to my numerous collaborators, and especially to Steve Kivelson and Eduardo Fradkin for many stimulating and illuminating conversations.
This work was supported by the Office of Basic Energy Sciences, Division of Materials Science and Engineering, U.S. Department of Energy (DOE), under Contract No. DE-AC02-98CH10886 and through the Center for Emergent Superconductivity, an Energy Frontier Research Center.
|
\section{Introduction}\label{sec:3-universal-priors}
In the study of universal induction, we consider an abstraction
of the world in the form of a binary string. Any sequence from
a finite set of possibilities can be expressed in this way, and
that is precisely what contemporary computers are capable of
analysing. An ``environment'' provides a measure of probability
to (possibly infinite) binary strings. Typically, the class
$\mathcal{M}$ of enumerable semimeasures is considered. Given
the equivalence between $\mathcal{M}$ and the set of monotone
Turing machines (Lemma \ref{lem:measures-TM}), this choice
reflects the expectation that the environment can be computed
by (or at least approximated by) a Turing machine.
Universal induction is an ideal Bayesian induction mechanism
assigning probabilities to possible continuations of a binary
string. In order to do this, a prior distribution, termed a
universal prior, is defined on binary strings. This prior has
the property that the Bayesian mechanism converges to the true
(generating) environment for \textit{any} environment $\mu$ in
$\mathcal{M}$, given sufficient evidence.
There are three popular ways of defining a universal prior in
the literature: Solomonoff's prior
\cite{Solomonoff:64,Zvonkin:70,Hutter:04uaibook},
\label{def:u-prior-Solomonofxf} as a universal mixture
\cite{Zvonkin:70,Hutter:04uaibook,Hutter:07uspx}, or
\label{def:u-prior-mixturex} a universally dominant semimeasure
\cite{Hutter:04uaibook,Hutter:07uspx}.
\label{def:u-prior-dominantx} Briefly, a universally dominant
semimeasure is one that dominates every other semimeasure in
$\mathcal{M}$ (Definition \ref{def:u-prior-dominant}), a
universal mixture is a mixture of all semimeasures in
$\mathcal{M}$ with non-zero coefficients (Definition
\ref{def:u-prior-mixture}), and a Solomonoff prior assigns the
probability that a (chosen) monotone universal Turing machine
outputs a string given random input (Definition
\ref{def:u-prior-Solomonof}). These and other relevant concepts
are defined in more detail in Section \ref{sec:definitions}.
Solomonoff's and the universal mixture constructions have been
known for many years and they are often used interchangeably in
textbooks and lecture notes. Their equivalence has been shown
in the sense that they dominate each other
\cite{Zvonkin:70,Hutter:04uaibook,Li:08}. We extend this result
in Section \ref{sec:UTM-is-mixture}, showing that they in fact
define exactly the same class of priors.
Further, it is trivial to see that both constructions produce
universally dominant semimeasures. The converse is, however,
not true. Universally dominant semimeasures are a larger class.
We provide a simple example to demonstrate this in Section
\ref{sec:dominant-is-not-universal}.
These results are relatively undemanding technically, however
given their fundamental nature,
that they have not to our knowledge been published to date, and
the relevance to Ray Solomonoff's famous work on universal
induction, we present them here.
The following diagram summarises these inclusion relations:
\begin{figure}
\begin{displaymath}
\xymatrix{
&\text{Universally Dominant} \ar@/_/@{<-}[ddl]_{Lemma \ref{lem:u-mix-is-dominant}} \ar@/^/@{.x}[ddl]^{Theorem\; \ref{thm:dom-not-mixture}}&\\
&&\\
\text{Universal Mixture}\ar@/_/@{<->}[rr]^{Theorem\; \ref{thm:UTM-eq-mixture}}
&& \text{Solomonoff Prior} \ar@/_/[uul]_{Corollary\; \ref{corol:UTM-is-dominant}}
}
\end{displaymath}
\caption{}
\end{figure}
\section{Definitions}\label{sec:definitions}
We represent the set of finite/infinite binary strings as
$\mathbb{B}^*$ and $\mathbb{B}^\infty$ respectively. $\epsilon$
denotes the empty string, $xb$ the concatenation of strings $x$
and $b$, $\ell(x)$ the length of a string $x$. A cylinder set,
the set of all infinite binary strings which start with some
$x\in\mathbb{B}^*$ is denoted $\Gamma_x$.
A string $x$ is said to be a prefix of a string $y$ if $y=xz$
for some string $z$. We write $x\sqsubseteq y$ or $x\sqsubset
y$ if $x$ is a proper substring of $y$ (ie: $z\ne\epsilon$). We
denote the maximal prefix-free subset of a set of finite
strings $\mathcal{P}$ by $\lfloor \mathcal{P} \rfloor$. It can
be obtained by successively removing elements that have a
prefix in $\mathcal{P}$. The uniform measure of a set of
strings is denoted $| \mathcal{P}
|:=\sum_{p\in\lfloor\mathcal{P}\rfloor}2^{-\ell(p)}$. This is
the area of continuations of elements of $\mathcal{P}$
considered as binary decimal numbers.
There have been several definitions of monotone Turing machines
in the literature \cite{Li:08}, however we choose that which is
now widely accepted
\cite{Solomonoff:64,Zvonkin:70,Hutter:04uaibook,Li:08}
and has the useful and intuitive property Lemma \ref{lem:measures-TM}.
\begin{definition}
A monotone Turing machine is a computer with binary (one-way)
input and output tapes, a bidirectional binary work tape (with
read/write heads as appropriate) and a finite state machine to
determine its actions given input and work tape values. The
input tape is read-only, the output tape is write-only.
\end{definition}
The definitions of a universal Turing machine in the literature
are somewhat varied or unclear. Monotone universal Turing
machines are relevant here for defining the Solomonoff prior.
In the algorithmic information theory literature, most authors
are concerned with the explicit construction of a single
reference universal machine
\cite{Hutter:04uaibook,Li:08,Solomonoff:64,Turing:36,Zvonkin:70}.
A more general definition is left to a relatively vague
statement along the lines of ``a Turing machine that can
emulate any other Turing machine''. The definition below
reflects the typical construction used and is often referred to
as \textit{universal by adjunction}
\cite{Downey:10book,Figueira:06}.
\begin{definition}[Monotone Universal Turing Machine]\label{def:UTM}
A monotone universal Turing machine is a monotone Turing
machine $U$ for which there exist:
\begin{enumerate}
\item an enumeration $\{T_i:i\in\mathbb{N}\}$ of all monotone Turing machines
\item a computable uniquely decodable self-delimiting code $I:\mathbb{N}\rightarrow\mathbb{B}^*$
\end{enumerate}
such that the programs for $U$ that produce output coincide
with the set $\{I(i)p:i\in\mathbb{N},\;p\in\mathbb{B}^*\}$ of
concatenations of $I(i)$ and $p$, and
\[
U(I(i)p) = T_i(p)\quad\forall\, i\in\mathbb{N}\;,\;p\in\mathbb{B}^*
\]
\end{definition}
A key concept in algorithmic information theory is the
assignment of probability to a string $x$ as the probability
that some monotone Turing machine produces output beginning
with $x$ given unbiased coin flip input. This approach was used
by Solomonoff to construct a universal prior
\cite{Solomonoff:64}. To better understand the properties of
such a function, we will need the concepts of enumerability
and semimeasures:
\begin{definition}
A function or number $\phi$ is said to be
\textbf{\emph{enumerable}} or \textbf{\emph{lower
semicomputable}} (these terms are synonymous) if it can be
approximated from below (pointwise) by a monotone increasing
set $\{\phi_i:i\in\mathbb{N}\}$ of finitely computable
functions/numbers, all calculable by a single Turing machine.
We write $\phi_i\nearrow\phi$. Finitely computable
functions/numbers can be computed in finite time by a Turing
machine.
\end{definition}
\begin{definition}
A \textbf{\emph{semimeasure}} is a ``defective'' probability
measure on the $\sigma$-algebra generated by cylinder sets in
$\mathbb{B}^\infty$. We write $\mu(x)$ for $x\in\mathbb{B}^*$
as shorthand for $\mu(\Gamma_x)$. A probability measure must
satisfy $\mu(\epsilon)=1$,
$\mu(x)=\sum_{b\in\mathbb{B}}\mu(xb)$. A semimeasure allows a
probability ``gap'': $\mu(\epsilon)\le1$ and
$\mu(x)\ge\sum_{b\in\mathbb{B}}\mu(xb)$. $\mathcal{M}$ denotes
the set of all enumerable semimeasures.
\end{definition}
The following definition explicates the relationship between
monotone Turing machines and enumerable semimeasures.
\begin{definition}[Solomonoff semimeasure]
\label{def:lambda_T}
For each monotone Turing machine $T$ we associate a semimeasure
\[
\lambda_T(x) := \sum_{\lfloor p:T(p)=x*\rfloor}2^{-\ell(p)} = |T^{-1}(x*)|
\]
where $\lfloor \mathcal{P} \rfloor$ indicates the maximal
prefix-free subset of a set of finite strings $\mathcal{P}$,
$T(p)=x*$ indicates that $x$ is a prefix of (or equal to)
$T(p)$ and $\ell(p)$ is the length of $p$.
If there are no such programs, we set $\lambda_T(x):=0$. [See
\cite{Li:08} definition 4.5.4]
\end{definition}
Note that this is the probability that $T$ outputs a string
starting with $x$ given unbiased coin flip input. To see this,
consider the uniform measure given by
$\lambda(\Gamma_p):=2^{-\ell(p)}$. This is the probability of
obtaining $p$ from unbiased coin flips. $\lambda_T(x)$ is the
uniform measure of the set of programs for $T$ that produce
output starting with $x$, ie: the probability of obtaining one
of those programs from unbiased coin flips. Note also that,
since $T$ is monotone, this set consists of a union of disjoint
cylinder sets $\{\Gamma_p:p\in\lfloor q:T(q)=x*\rfloor\}$. By
dovetailing a search for such programs and an lower
approximation of the uniform measure $\lambda$, we can see that
$\lambda_T$ is enumerable. See Definition 4.5.4 (p.299) and
Lemma 4.5.5 (p.300) in \cite{Li:08}.
An important lemma in this discussion establishes the
equivalence between the set of all monotone Turing machines and
the set $\mathcal{M}$ of all enumerable semimeasures. It is
equivalent to Theorem 4.5.2 in \cite{Li:08} (page 301) with a
small correction: $\lambda_T(\epsilon)=1$ for any $T$ by
construction, but $\mu(\epsilon)$ may not be $1$, so this case
must be excluded.
\begin{lemma}\label{lem:measures-TM}
A semimeasure $\mu$ is lower semicomputable if and only if
there is a monotone Turing machine $T$ such that
$\mu=\lambda_T$ except on $\Gamma_\epsilon \equiv
\mathbb{B}^\infty$ and $\mu(\epsilon)$ is lower semicomputable.
\end{lemma}
We are now equipped to formally define the 3 formulations for a
universal prior:
\begin{definition}[Solomonoff prior]
\label{def:u-prior-Solomonof}
The Solomonoff prior for a given universal monotone Turing machine $U$ is
\[
M:=\lambda_U
\]
The class of all Solomonoff priors we denote $\mathcal{U}_M$.
\end{definition}
\begin{definition}[Universal mixture]\label{def:u-prior-mixture}
A universal mixture is a mixture $\xi$ with non-zero positive
weights over an enumeration $\{\nu_i:i\in\mathbb{N},
\nu_i\in\mathcal{M}\}$ of all enumerable semimeasures
$\mathcal{M}$:
\[
\xi = \sum_{i\in\mathbb{N}}w_i\nu_i\quad:\quad \mathbb{R}\ni w_i>0\;,\;\sum_{i\in\mathbb{N}}w_i\le1
\]
We require the weights $w_{()}$ to be a lower semicomputable
function. The mixture $\xi$ is then itself an enumerable
semimeasure, i.e. $\xi\in\mathcal{M}$. The class of all
universal mixtures we denote $\mathcal{U}_\xi$.
\end{definition}
\begin{definition}[Universally dominant semimeasure]\label{def:u-prior-dominant}
A universally dominant semimeasure is an enumerable semimeasure
$\delta$ for which there exists a real number $c_\mu>0$ for
each enumerable semimeasure $\mu$ satisfying:
\[
\delta(x) \ge c_\mu\mu(x)\quad\forall x\in\mathbb{B}^*
\]
The class of all universally dominant semimeasures we denote
$\mathcal{U}_\delta$.
\end{definition}
Dominance implies absolute continuity: Every enumerable
semimeasure is absolutely continuous with respect to a
universally dominant enumerable semimeasure. The converse
(absolute continuity implies dominance) is however not true.
\section{Equivalence between Solomonoff priors and universal mixtures}\label{sec:UTM-is-mixture}
We show here that every Solomonoff prior $M\in\mathcal{U}_M$
can be expressed as a universal mixture (i.e.:
$M\in\mathcal{U}_\xi$) and vice versa. In other words the class
of Solomonoff priors and the class of universal mixtures are
identical: $\mathcal{U}_M=\mathcal{U}_\xi$.
Previously, it was known
\cite{Zvonkin:70,Hutter:04uaibook,Li:08} that a Solomonoff
prior $M$ and a universal mixture $\xi$ are equivalent up to
multiplicative constants
\begin{align}
M(x) &\le c_1\xi(x) &\forall x\in\mathbb{B}^* \notag\\
\xi(x) &\le c_2M(x) &\forall x\in\mathbb{B}^* \notag
\end{align}
The result we present is stronger, stating that the two classes
are exactly identical. Again we exclude the case $x=\epsilon$
as $M(\epsilon)$ is always one for a Solomonoff prior, but
$\xi(\epsilon)$ is never one for a universal mixture $\xi$ (as
there are $\mu\in\mathcal{M}$ with $\mu(\epsilon)<1$).
\begin{lemma} \label{thm:UTM-is-mixture}
For any monotone universal Turing machine $U$ the associated
\linebreak Solomonoff prior $M$ can be expressed as a universal
mixture. i.e. there exists an enumeration
$\{\nu_i\}_{i=1}^\infty$ of the set of enumerable semimeasures
$\mathcal{M}$ and computable function
$w_{()}:\mathbb{N}\rightarrow\mathbb{R}$ such that
\[
M(x)=\sum_{i\in\mathbb{N}} w_i\nu_i(x)\quad\forall x\in\mathbb{B}^*\backslash\epsilon
\]
with $\sum_{i\in\mathbb{N}} w_i\le 1$ and $w_i>0\;\forall
i\in\mathbb{N}$. In other words the class of Solomonoff priors
is a subset of the class of universal mixtures:
$\mathcal{U}_M\subseteq\mathcal{U}_\xi$.
\end{lemma}
\begin{proof}
We note that all programs that produce output from $U$ are
uniquely of the form $q=I(i)p$. This allows us to split the sum
in (\ref{eqn:split-sum}) below.
\begin{align}
M(x) &= \sum_{\lfloor q:U(q)=x*\rfloor}2^{-\ell(q)} &\notag\\
&=\sum_{i\in\mathbb{N}}\sum_{\lfloor p:U(I(i)p)=x*\rfloor}2^{-\ell(I(i)p)} &\label{eqn:split-sum} \\
&=\sum_{i\in\mathbb{N}}2^{-l(I(i))}\sum_{\lfloor p:T_i(p)=x*\rfloor}2^{-\ell(p)} & \label{eqn:UTM-is-mix-take-prefix} \notag\\
&=\sum_{i\in\mathbb{N}}2^{-l(I(i))}\lambda_{T_i}(x) &\notag
\end{align}
Clearly $2^{-l(I(i))}>0$ and is a computable function of $i$.
Since $I$ is a self-delimiting code it must be prefix free, and
so satisfy Kraft's inequality:
\begin{equation}
\sum_{i\in\mathbb{N}}2^{-l(I(i))} \le 1 \notag
\end{equation}
Lemma \ref{lem:measures-TM} tells us that the $\lambda_{T_i}$
cover every enumerable semimeasure if $\epsilon$ is excluded
from their domain, which shows that
$\sum_{i\in\mathbb{N}}2^{-l(I(i))}\lambda_{T_i}(x)$ is a
universal mixture. This completes the proof.
\end{proof}
\begin{corollary} \label{corol:UTM-is-dominant}
\cite{Zvonkin:70} The Solomonoff prior $M$ for a universal
monotone Turing machine $U$ is universally dominant. Thus, the
class of Solomonoff priors is a subset of the class of
universally dominant lower semicomputable semimeasures:
$\mathcal{U}_M\subseteq\mathcal{U}_\delta$.
\end{corollary}
\begin{proof}
From Lemma \ref{thm:UTM-is-mixture} we have for each
$\nu\in\mathcal{M}$ there exists $j\in\mathbb{N}$ with
$\nu=\lambda_{T_j}$ and for all $x\in\mathbb{B}^*$:
\begin{align*}
M(x) &= \sum_{i\in\mathbb{N}}2^{-l(I(i))}\lambda_{T_i}(x) \\
&\ge 2^{-l(I(j))}\nu(x)
\end{align*}
as required.
\end{proof}
\begin{lemma}\label{lem:u-mix-is-dominant}
Every universal mixture $\xi$ is universally dominant. Thus,
the class of universal mixtures is a subset of the class of
universally dominant lower semicomputable semimeasures:
$\mathcal{U}_\xi\subseteq\mathcal{U}_\delta$.
\end{lemma}
\begin{proof}
This follows from a similar argument to that in Corollary
\ref{corol:UTM-is-dominant}.
\end{proof}
\begin{lemma} \label{thm:mixture-is-UTM}
For every universal mixture $\xi$ there exists a universal
monotone Turing machine and associated Solomonoff prior $M$
such that
\[
\xi(x)=M(x)\quad\forall x\in\mathbb{B}^*\backslash\epsilon
\]
In other words the class of universal mixtures is a subset of
the class of Solomonoff priors:
$\mathcal{U}_\xi\subseteq\mathcal{U}_M$.
\end{lemma}
\begin{proof}
First note that by Lemma \ref{lem:measures-TM} we can find (by
dovetailing possible repetitions of some indicies) parallel
enumerations $\{\nu_i\}_{i\in\mathbb{N}}$ of $\mathcal{M}$ and
$\{T_i=\lambda_{\nu_i}\}_{i\in\mathbb{N}}$ of all monotone
Turing machines, and computable weight function $w_{()}$ with
\[
\xi = \sum_{i\in\mathbb{N}} w_i\nu_i \quad , \quad \sum_{i\in\mathbb{N}}w_i \le 1
\]
Take a computable index and lower approximation
$\phi(i,t)\nearrow w_i$:
\begin{align}
w_i &= \sum_t|\phi(i,t+1)-\phi(i,t)| \\
&= \sum_j 2^{-k_{ij}}\\
i,j&\mapsto k_{ij} \;\text{computable}
\end{align}
The K-C theorem
\cite{Levin:71,Schnorr:73,Chaitin:75,Downey:10book} says that
for any computable sequence of pairs $ \{k_{ij}\in\mathbb{N},\;
\tau_{ij} \in\mathbb{B}^*\}_{i,j\in\mathbb{N}}$ with $\sum
2^{-k_{ij}}\le 1$, there exists a prefix Turing machine $P$ and
strings $\{\sigma_{ij}\in\mathbb{B}^*\}$ such that
\begin{equation}
\ell(\sigma_{ij})=k_{ij}\;,\;P(\sigma_{ij})=\tau_{ij}
\end{equation}
Choosing distinct $\tau_{ij}$ and the existence of prefix
machine $P$ ensures that $\{\sigma_{ij}\}$ is prefix free. We
now define a monotone Turing machine $U$. For strings of the
form $\sigma_{ij}p$ for some $i,j$:
\begin{equation}
U(\sigma_{ij}p) := T_i(p)
\end{equation}
For strings not of this form, $U$ produces no output. $U$
inherits monotonicity from the $T_i$, and since
$\{T_i\}_{i\in\mathbb{N}}$ enumerates all monotone Turing
machines, $U$ is universal. The Solomonoff prior associated
with $U$ is then:
\begin{align}
\lambda_U(x) &= |U^{-1}(x*)| \\
&= \sum_{i,j}2^{-\ell(\sigma_{ij})}|T_i^{-1}(x*)| \\
&= \sum_i (\sum_j2^{-k_{ij}})\lambda_{T_i}(x) \\
&= \sum_i w_i \nu_i(x) \\
&= \xi(x)
\end{align}
\end{proof}
The main theorem for this section is now trivial:
\begin{theorem} \label{thm:UTM-eq-mixture}
The classes $\mathcal{U}_M$ of Solomonoff priors and
$\mathcal{U}_\xi$ of universal mixtures are exactly equivalent.
In other words, the two constructions define exactly the same
set of priors: $\mathcal{U}_M=\mathcal{U}_\xi$.
\end{theorem}
\begin{proof}
Follows directly from Lemma \ref{thm:UTM-is-mixture} and Lemma
\ref{thm:mixture-is-UTM}.
\end{proof}
\section{Not all universally dominant enumerable semimeasures are universal mixtures}\label{sec:dominant-is-not-universal}
In this section, we see that a universal mixture must have a
``gap'' in the semimeasure inequality greater than
$c\,2^{-K(\ell(x))}$ for some constant $c>0$ independent of
$x$, and that there are universally dominant enumerable
semimeasures that fail this requirement. This shows that not
all universally dominant enumerable semimeasures are universal
mixtures.
\begin{lemma}
\label{lem:u-mix-no-gaps} For every Solomonoff prior $M$ and
associated universal monotone Turing machine $U$, there exists
a real constant $c>0$ such that
\[
\frac{M(x)-M(x0)-M(x1)}{M(x)}\ge c\,2^{-K(\ell(x))}\quad\forall x\in\mathbb{B}^*
\]
where the Kolmogorov complexity $K(n)$ of an integer $n$ is the
length of the shortest prefix code for $n$.
\end{lemma}
\begin{proof}
First, note that $M(x)-M(x0)-M(x1)$ measures the set of
programs $U^{-1}(x)$ for which $U$ outputs $x$ and no more.
Consider the set
\[
\mathcal{P}:=\{ql'p\,|\,p\in \mathbb{B}^*,\,U(p)\sqsupseteq x\}
\]
where $l'$ is a shortest prefix code for $\ell(x)$ and $q$ is a
program such that $U(q{l}'p)$ executes $U(p)$ until $\ell(x)$
bits are output, then stops.
Now, for each $r=q{l}'p\in\mathcal{P}$ we have $U(r)=x$ since
$U(p)\sqsupseteq x$ and $q$ executes $U(p)$ until $\ell(x)$
bits are output. Thus $\mathcal{P}\subseteq U^{-1}(x)$ and
\begin{equation}
|\mathcal{P}|\le |U^{-1}(x)| \label{eqn:p-u-1x}
\end{equation}
Also $\mathcal{P}=q{l}'U^{-1}(x*):=\{s=q{l}'p\,|\,p\in U^{-1}(x*)\}$, and so
\begin{equation}
|\mathcal{P}|=2^{-\ell(q{l}')}|U^{-1}(x*)| \label{eqn:p-u-1x*}
\end{equation}
combining (\ref{eqn:p-u-1x}) and (\ref{eqn:p-u-1x*}) and noting
that $M(x)-M(x0)-M(x1)=|U^{-1}(x)|$ and $M(x)=|U^{-1}(x*)|$ we
obtain
\begin{align}
M(x)-M(x0)-M(x1) &= |U^{-1}(x)| \notag \\
&\ge |\mathcal{P}| \notag\\
&= 2^{-\ell(q{l}')}|U^{-1}(x*)| \notag \\
&= 2^{-\ell(q)}2^{-K(\ell(x))}M(x) \notag
\end{align}
Setting $c:=2^{-\ell(q)}$ this proves the result.
\end{proof}
\begin{theorem} \label{thm:dom-not-mixture}
Not all universally dominant enumerable semimeasures are
universal mixtures: $\mathcal{U}_\xi\subset\mathcal{U}_\delta$
\end{theorem}
\begin{proof}
Take some universally dominant semimeasure $\delta$, then define
$
\delta'(\epsilon):= 1,\;
\delta'(0)=
\delta'(1):=\frac{1}{2},\;
\delta'(bx):=\frac{1}{2}\delta(bx)$ for $b\in\mathbb{B}$, $x\in\mathbb{B}^*\backslash\epsilon
$. $\delta'$ is clearly a universally dominant enumerable
semimeasure with $\delta'(0)+\delta'(1)=\delta'(\epsilon)$, and
by Lemma \ref{lem:u-mix-no-gaps} it is not a universal mixture.
\end{proof}
\section{Conclusions}
One of Solomonoff's more famous contributions is the invention
of a theoretically ideal universal induction mechanism. The
universal prior used in this mechanism can be
defined/constructed in several ways.
We clarify the relationships between three different
definitions of universal priors, namely universal mixtures,
Solomonoff priors and universally dominant semimeasures. We
show that the class of universal mixtures and the class of
Solomonoff priors are exactly the same while the class of
universally dominant lower semicomputable semimeasures is a
strictly larger set.
We have identified some aspects of the discrepancy between
Solomonoff priors/universal mixtures and universally dominant
lower semicomputable semimeasures, however a clearer
understanding and characterisation would be of interest.
Since universal dominance is all that is needed to prove
convergence for universal induction
\cite{Hutter:04uaibook,Solomonoff:78} it is interesting to ask
whether the extra properties of the smaller class of Solomonoff
priors have any positive consequences for universal induction.
\subsubsection*{Acknowledgements.}
We would like to acknowledge the contribution of an anonymous
reviewer to a more elegant presentation of the proof of Lemma
\ref{thm:mixture-is-UTM}. This work was supported by ARC grant
DP0988049.
\begin{small}
|
\section{Introduction}
Basic arguments suggest that the angular frequency $\Omega(r)$ of an
accretion disk surrounding a weakly magnetized star must attain a
maximum value, $\Omega_{\rm max} \equiv \Omega(r_\mathrm{b})$, and decrease
inward \citep[or at least remain constant;
see][]{2004ApJ...613..506M} to match the angular frequency of the
star at the stellar radius $\Omega_\star(r_\star)$; see, e.g.,
\cite{2002apa..book.....F, 2009apsf.book.....H, 2010apf..book.....A}.
The inner disk region, where $r < r_\mathrm{b}$ and $d\Omega/dr \ge 0$, is
referred to as the accretion disk \emph{boundary layer}.
Standard accretion disk theory \citep{1973A&A....24..337S} predicts
that half of the energy released in the accretion process takes place
in this region, estimated to be a fraction of the stellar radius.
The spectrum of the radiated energy depends on the detailed properties
of this layer \citep{1993Natur.362..820N, 1995ApJ...442..337P,
1996ApJ...473..422P, 2001ApJ...547..355P}; thus understanding the
various processes that determine its properties \citep[see,
e.g.,][]{2004ApJ...610..977P, 2009ApJ...702.1536B,
2010AstL...36..848I} is of fundamental importance.
Most detailed calculations for determining the structure of the
boundary layer rely on effective models for turbulent angular momentum
transport. These models are usually built as a turbulent version of the Newtonian
viscous stress between fluid layers in a differentially rotating
laminar flow \citep{1959flme.book.....L}, and thus assume a linear
relationship between the stress and the angular frequency gradient
\citep{1974MNRAS.168..603L}.
This assumption, however, seems at odds with the properties of
magnetohydrodynamic (MHD) turbulence revealed by numerical simulations
of accretion disks \citep{1996MNRAS.281L..21A, 2002MNRAS.330..895A,
2002ApJ...571..413S, 2008MNRAS.383..683P} which show that angular
momentum transport is inefficient in regions of the disk where
$d\Omega/dr > 0$, which are stable to the standard magnetorotational
instability \citep[MRI; see][]{1991ApJ...376..214B,
1998RvMP...70....1B}.
Motivated by the need of a deeper understanding of the behavior of an
MHD fluid in a differentially rotating background that deviates from a
Keplerian profile, we study the dynamics of MHD waves in
configurations that are stable to the standard MRI.
Employing the shearing-sheet framework, we show that transient
amplification of shearing MHD waves can generate magnetic energy
without leading to a substantial generation of hydromagnetic stresses.
We discuss the implications of these findings.
\section{Assumptions and Local Model for MHD Disk}
We focus our attention on a subsonic, weakly magnetized fluid for
which the ram pressure and magnetic pressure remain small compared to
their thermal counterpart.
As a first approximation, we thus consider a differentially rotating
fluid with angular frequency $\V{\Omega} = \Omega(r) \B{z}$ and
constant background density $\rho_0$.
This is a reasonable assumption in light of the results presented by
\cite{2002MNRAS.330..895A} and \cite{2002ApJ...571..413S}, who carried
out numerical simulations of boundary layers of unstratified accretion
disks and found that the density fluctuations throughout the simulations
are in general quite small.
We work in the framework of the shearing-sheet approximation
\citep{Hill1878, 1978ApJ...222..850G, 1987MNRAS.228....1N,
1995ApJ...440..742H}, where the equations describing an
incompressible MHD fluid in a corotating frame are given by
\begin{align}
\partial_t \V{v} + (\V{v}\cdot\nabla)\V{v}
&= - 2\Omega_0\B{z}\times\V{v} + 2 q\Omega_0^2 x\B{x}
- \smash{\frac{\nabla P}{\rho_0}} \nonumber\\
&+ \frac{(\V{B}\cdot\nabla)\V{B}}{4\pi\rho_0} + \nu\nabla^2\V{v}
\,, \label{eq:v}\\
\partial_t \V{B} + (\V{v}\cdot\nabla)\V{B}
&= (\V{B}\cdot\nabla)\V{v} + \eta\nabla^2\V{B} \,. \label{eq:B}
\end{align}
Here, $\V{v}(\V{x},t)$ and $\V{B}(\V{x},t)$, with $\nabla\cdot\V{v} =
\nabla\cdot\V{B} = 0$, stand for the velocity and magnetic fields;
$\nu$ and $\eta$ denote the kinematic viscosity and resistivity; and
$\Omega_0 \equiv \Omega(r_0)$ is the corotating angular frequency at a
fiducial radius $r_0$.
The first and second terms on the right hand side of
equation~(\ref{eq:v}) correspond to the Coriolis and tidal forces,
respectively. The local pressure $P$ can be found by the
divergence-less condition of the velocity field.
Recalling that the local density $\rho_0$ is assumed to be constant,
and in order to simplify notation, hereafter we redefine the symbols
denoting the pressure and magnetic field in such a way that and
$P/\rho_0 \rightarrow P$ and $\V{B}/(4\pi\rho_0)^{1/2} \rightarrow
\V{B}$.
We decompose the flow into mean and fluctuations as
\begin{align}
\V{v}(\V{x}, t) &\equiv \V{U}_1(x) + \V{u}(\V{x}, t), \label{eq:v_decomp}\\
\V{B}(\V{x}, t) &\equiv \V{B}_0(t) + \V{b}(\V{x}, t). \label{eq:B_decomp}
\end{align}
The leading order background velocity is $\V{U}_1(x) \equiv
-q\,\Omega_0 x \B{y}$, where the shear parameter $q$ is given by
\begin{align}
q \equiv \left.-\frac{d\ln \Omega}{d\ln r}\right|_{r_0} \,.
\end{align}
The homogeneous background magnetic field is in general a function of
time and it evolves according to the induction equation~(\ref{eq:B}),
i.e., $\partial_t\V{B}_0 = - q\,\Omega_0 B_{0x} \B{y}$.
The substitution of equations~(\ref{eq:v_decomp}) and
(\ref{eq:B_decomp}) into (\ref{eq:v}) and (\ref{eq:B}) leads to a
non-linear system for the dynamical evolution of the perturbations
$\V{u}(\V{x}, t)$ and $\V{b}(\V{x}, t)$.
However, as pointed out in \citet{1994ApJ...432..213G}, all the
non-linear terms in the resulting equations vanish identically if we
consider the evolution of a single Fourier mode\footnote{Formally
speaking we consider two modes, with wavenumbers $\V{k}$ and $-\V{k}$.
However, because the functions under consideration are real, the
Fourier coefficients satisfy $\f{f}_{-\V{k}} = \f{f}_{\V{k}}^*$.}.
In this case we obtain
\begin{align}
\mathcal{D}_t\V{u} - q\,\Omega_0 u_x \B{y}
&= \V{B}_0\cdot\nabla\V{b} + \nu \nabla^2\V{u}
- 2 \Omega_0 \B{z}\times\V{u} - \nabla P, \label{eq:u}\\
\mathcal{D}_t\V{b} + q\,\Omega_0 b_x \B{y}
&= \V{B}_0\cdot\nabla\V{u} + \eta\nabla^2\V{b}. \label{eq:b}
\end{align}
The ``semi-Lagrangian'' time derivative $\mathcal{D}_t \equiv
\partial_t + \V{U}_1\cdot\nabla$ accounts for advection by the
shearing background.
The shearing component in the Coriolis term cancels out the tidal
force.
We remark that equations~(\ref{eq:u}) and (\ref{eq:b}) are \emph{not}
just linearized equations, they remain valid even if the amplitude of
the perturbations is not small compared to the background values, and
they are exact as long as a single Fourier mode is considered.
Under these conditions, it is sensible to study the evolution of
$\V{u}(\V{x}, t)$ and $\V{b}(\V{x}, t)$ for a long time.
In order to solve equations~(\ref{eq:u}) and (\ref{eq:b}), it is
convenient to work in Fourier space.
The $x$-dependence of the ``semi-Lagrangian'' time derivative can be
removed by employing a shearing coordinate system $(x', y', z', t')
\equiv (x, y + q\,\Omega_0 x t, z, t)$ in which $\mathcal{D}_t =
\partial_{t'}$ \citep{1965MNRAS.130..125G}.
A single mode with a fixed ``shearing'' wavenumber $\V{k}'$ is thus
given by
\begin{align}
\V{u}(\V{x},t) &= 2\mathrm{Re}
\left[\F{u}_{\V{k}'}(t)\exp(i\V{k}'\cdot\V{x}')\right],\label{eq:u_fourier}\\
\V{b}(\V{x},t) &= 2\mathrm{Re}
\left[\F{b}_{\V{k}'}(t)\exp(i\V{k}'\cdot\V{x}')\right],\label{eq:b_fourier}
\end{align}
where $\V{k}'\cdot\V{x}' = \V{k}(t)\cdot\V{x} = (k_x' + q\,\Omega_0 t
k_y') x + k_y' y + k_z' z$.
Substituting the ansatz~(\ref{eq:u_fourier}) and (\ref{eq:b_fourier})
into equations~(\ref{eq:u}) and (\ref{eq:b}) leads a set of equations
for the Fourier amplitudes
\begin{align}
d_t\F{u} - q\,\Omega_0 \f{u}_x \B{y} &= i\omega_\mathrm{A}\F{b} - \nu k^2\F{u}
- 2 \Omega_0 \B{z}\times\F{u} - i\V{k} \f{P},
\label{eq:uk}\\
d_t\F{b} + q\,\Omega_0 \f{b}_x \B{y} &= i\omega_\mathrm{A}\F{u} - \eta k^2\F{b}.
\label{eq:bk}
\end{align}
Here, we have replaced $\partial_{t '}$ by $d_t$ and omitted the
subscripts $\V{k}'$ in the Fourier coefficients in order to
simplify the notation. We have also introduced the (time-independent) Alfv\'en frequency
$\omega_\mathrm{A} \equiv \V{B}_0(t)\cdot\V{k}(t)$ (see also \citealt{1992ApJ...400..610B}).
The pressure term can be eliminated from equation~(\ref{eq:uk}) using
the solenoidal character of the velocity field, which implies $d_t
i\V{k} = i\V{k} d_t + iq\,\Omega_0 k_y\B{x}$.
This leads to
\begin{align}
\f{P} = -\frac{2i\Omega_0}{k^2}\Big[(q-1)k_y\f{u}_x + k_x\f{u}_y\Big],
\label{eq:pressure_term}
\end{align}
which is independent of $\f{u}_z$.
Because we are interested in the transport of angular momentum along
the radial direction, this decoupling allows us to solve separately
for the $x$- and $y$-components.
\section{Dynamical Evolution of MRI-stable Modes}
\begin{figure*}
\includegraphics[width=\columnwidth,trim=16 8 8 16]{f1a.eps}
\hfill
\includegraphics[width=\columnwidth,trim=16 8 8 16]{f1b.eps}
\caption{The Fourier amplitudes for the magnetic field components
$\hat{b}_x$ (left) and $\hat{b}_y$ (right).
The thick solid lines are the numerical solutions obtained by
solving equation~(\ref{eq:2nd}) with $q= -3/2$ and
$\omega_\mathrm{A} = \Omega_0$ and using the divergence-less conditions~(\ref{eq:divless}).
The asymptotic solutions for early and late times are both given
by the analytical approximation~(\ref{app:bx}) and (\ref{app:by})
but with different integration constants.
The dotted lines correspond to $C_- = 1$ and $S_- = 0$, which are
used to specify the initial conditions at $\omega\,\tau = -20$.
The dashed lines correspond to $C_+ = 0.285$ and $S_+ = -1.896$,
which are obtained by matching the analytic and numerical results
at late times, e.g., $\omega\,\tau = 20$.
Note that, for the sake of clarity, we truncate the analytical
solutions in the right panel at $\omega\,\tau = 2$ and $-2$.}
\label{fig:uxbx}
\end{figure*}
The dynamical evolution of the modes with $\V{k} \equiv k_z\B{z}$ is
quite simple; they grow exponentially if $k_z^2 \omega_\mathrm{A}^2 \le 2 q
\Omega_0^2$ \citep{1992ApJ...392..662B, 2006MNRAS.372..183P}.
Thus, a Keplerian disk (with $q=3/2$) can exhibit exponential growth but a shear
profile with $q < 0$ only supports stable oscillations.
In order to isolate the interesting dynamics that could arise from
modes that are not associated with the MRI, we thus focus on modes
with $k_z = 0$.
Taking the curl of the momentum equation, it is easy to verify that
the Coriolis term does not play a role in the equation for the
vorticity, and hence it has no effect on the dynamics of the system
\citep{2007ApJ...670..789L}.
This shows explicitly that the standard MRI is absent in our analysis.
We choose the origin of time so that $k_x(t)$ is initially zero.
In other words, we use $k_x(t)$ to define our time coordinate,
\begin{align}
\tau \equiv k_x(t)/k_y \equiv q\,\Omega_0 t \,,
\end{align}
so the divergence-less conditions become
\begin{align}
\tau\,\f{u}_x + \f{u}_y = \tau\,\f{b}_x + \f{b}_y = 0. \label{eq:divless}
\end{align}
Assuming $\nu, \eta \ll \omega_\mathrm{A}/k^2$, we can work in the ideal limit and
neglect viscosity and resistivity.
The $x$-components of the MHD equations then become
\begin{align}
d_\tau
\begin{pmatrix}\f{u}_x \\ \f{b}_x\end{pmatrix} =
\begin{pmatrix}-\Gamma & i\omega \\ i\omega & 0 \\\end{pmatrix}
\begin{pmatrix}\f{u}_x \\ \f{b}_x\end{pmatrix}, \label{eq:uxbx}
\end{align}
where all the temporal dependence is contained in the factor
\begin{align}
\Gamma(\tau) \equiv 2 \tau / (\tau^2 + 1), \label{def:Gamma}
\end{align}
and
\begin{align}
\omega \equiv \omega_\mathrm{A} / q\,\Omega_0
\end{align}
is the dimensionless Alfv\'en frequency.
The linear system (\ref{eq:uxbx}) can be recast into one equation as
\begin{align}
d_\tau^2 \f{b}_x + \Gamma(\tau)\,d_\tau \f{b}_x + \omega^2 \f{b}_x = 0,
\label{eq:2nd}
\end{align}
where the dependence on the parameters $q$, $\Omega_0$, and $\omega_\mathrm{A}$ is
\emph{only} through the combination $(\omega_\mathrm{A} / q\,\Omega_0)^2$.
This second order differential equation for $\f{b}_x$ is identical to
equation (2.20) in \citet{1992ApJ...400..610B} when $k_z=0$.
In this case, the perturbations in the $z$-coordinate decouple from
the perturbations in the perpendicular direction, see also their
equation (2.19).
Unfortunately, equation~(\ref{eq:2nd}) does not have an analytical
solution.
However, if we consider the limit $\tau^2 \gg 1$, it reduces to a
spherical Bessel equation, which posseses as solutions
\begin{align}
\f{u}_x &= S j_1(\omega\,\tau) + C y_1(\omega\,\tau), \nonumber\\
&= - \frac{C + S \omega\,\tau}{\omega^2 \tau^2}\cos(\omega\,\tau)
+ \frac{S - C \omega\,\tau}{\omega^2 \tau^2}\sin(\omega\,\tau),
\label{app:ux}\\
\f{b}_x &= -i S j_0(\omega\,\tau) - i C y_0(\omega\,\tau) \nonumber\\
&= \frac{iC}{\omega\,\tau}\cos(\omega\,\tau)
- \frac{iS}{\omega\,\tau}\sin(\omega\,\tau),
\label{app:bx}
\end{align}
where $j_n(x)$ and $y_n(x)$, with $n=0,1$, are spherical Bessel
functions of the first and second kind, respectively; and $S$ and $C$
are complex constants determined by the initial conditions.
Using the divergence-less conditions in equation~(\ref{eq:divless}), the
$y$-components are simply
\begin{align}
\f{u}_y
&= - \tau\left[S j_1(\omega\,\tau) + C y_1(\omega\,\tau)\right]\nonumber\\
&= \frac{C + S\omega\tau}{\omega^2\tau}\cos(\omega\,\tau)
- \frac{S - C\omega\tau}{\omega^2\tau}\sin(\omega\,\tau), \label{app:uy}\\
\f{b}_y
&=\,\,i\tau\left[S j_0(\omega\,\tau) + C y_0(\omega\,\tau)\right]\nonumber\\
&= - \frac{iC}{\omega}\cos(\omega\,\tau)
+ \frac{iS}{\omega}\sin(\omega\,\tau). \label{app:by}
\end{align}
Because the pressure in equation~(\ref{eq:pressure_term}) is
independent of $\f{u}_z$, the exact solutions (for all time) are
Alfv\'en waves,
\begin{align}
\f{u}_z &=\,C'\cos(\omega\,\tau) +\;S'\sin(\omega\,\tau), \label{sol:uz}\\
\f{b}_z &= iC'\sin(\omega\,\tau) - iS'\cos(\omega\,\tau), \label{sol:bz}
\end{align}
where $C'$ and $S'$ are some other complex constants. We note that,
although the direction of the dimensionless time $\tau$ depends on the
sign of the shear parameter $q$, the combination $\omega\,\tau \equiv
\omega_\mathrm{A} t$ is insensitive to it.
Hence, given the same initial conditions, $\f{u}_x$ and $\f{b}_x$ are
symmetric, while $\f{u}_y$ and $\f{b}_y$ are anti-symmetric, in the
shear parameter $q$.
We demonstrate the accuracy of these analytical approximations in
Figure~\ref{fig:uxbx}, which shows both the numerical and analytical
solutions for $\mathrm{Im}[\f{b}_x]$ and $\mathrm{Im}[\f{b}_y]$ with
$q = -3/2$ and $\omega_\mathrm{A} = \Omega_0$ as an example.
The initial conditions are set at $\omega\,\tau = \omega_\mathrm{A} t = -20$ by
choosing $C_- = 1$ and $S_- = 0$.
The numerical solutions, shown with thick solid lines, result from
integrating equation~(\ref{eq:2nd}) with the
definition~(\ref{def:Gamma}) and using the divergence-less
conditions~(\ref{eq:divless}).
The dotted lines in the two panels are obtained by setting $C = C_-$
and $S = S_-$ in the analytical approximations~(\ref{app:bx}) and
(\ref{app:by}).
These solutions are indistinguishable for $\tau \lesssim -1$.
As expected, the approximations break down for $\tau \simeq 0$.
This is precisely where the numerical solutions change their
amplitudes significantly.
The analytical expressions~(\ref{app:bx}) and (\ref{app:by}) are again
in excellent agreement with the numerical solutions for $\tau \gtrsim
1$, provided that their amplitudes are given by $C = C_+ = 0.285$ and
$S = S_+ = -1.896$.
These constants are found by requiring that both the numerical and
analytical solutions match for $\omega\,\tau = \omega_\mathrm{A} t \gg 1$ (in
practice we set $\omega\,\tau=20$).
Even though our analytical approach cannot predict the change in
amplitude close to $\tau \approx 0$, the solutions that we obtain
are a very good approximation to the numerical results as long as $\tau^2 > 1$.
We could in principle obtain the coefficients $C_+$ and $S_+$ by an
asymptotic matching technique similar to the one employed in
\citet{2009MNRAS.397...52H}.
However, the analytical solution near $\tau \simeq 0$ contains special
functions that are too complicated to be useful.
More importantly, as we show below, the most interesting features of the
solutions are independent of the precise values of these constants.
\begin{figure}
\includegraphics[width=\columnwidth,trim=16 8 8 16]{f2.eps}
\caption{The thick and thin solid lines correspond, respectively, to
the total energy $E(t)$ and total stress $T_{xy}(t)$, calculated
using the Fourier amplitudes $\f{u}_x$, $\f{u}_y$, $\f{b}_x$, and
$\f{b}_y$ obtained numerically.
The dotted and dashed lines show the analytical approximation for
the energy, equation~(\ref{eq:E}), using the two sets of constants
described in the caption of Figure~\ref{fig:uxbx}.
The $x$-components of the velocity and magnetic fields are
symmetric in $q$ but the $y$-components are anti-symmetric.
Therefore, changing the sign of the shear parameter $q$ changes the
sign of the stress $T_{xy}$ but not of the energy $E$.}
\label{fig:se}
\end{figure}
\section{Late Time Stress and Energy}
Given the solutions~(\ref{app:ux}) -- (\ref{sol:bz}) for the Fourier
amplitudes, we obtain the (mean) total stress $\smash{T_{xy}} \equiv
\langle \smash{u_x u_y} - \smash{b_x b_y}\rangle$ and (mean) energy
density $E\equiv \langle \smash{u^2 + b^2} \rangle/2$ of the
fluctuating fields\footnote{We do not include the stress and energy
generated by the time-dependent mean field $\V{B}_0(t)$ here.
This contribution depends on the initial magnetic field.}, where the
brackets stand for the spatial average, see, e.g.,
\citet{2006MNRAS.372..183P}.
Because these solutions are only valid for early/late times, we can
approximate the total stress and energy density up to first order in
$1/\omega\,\tau$ as
\begin{align}
\!T_{xy} \!\approx\!-\frac{2}{\omega^2\tau}
&\Big[(|S|^2 - |C|^2)\cos(2\omega\,\tau)
+ ( S^*C + SC^*)\sin(2\omega\,\tau)\Big],
\label{eq:Txy}\\
E \!\approx\! -\smash{\frac{1}{\omega^3\tau}}
&\Big[(|S|^2 - |C|^2)\sin(2\omega\,\tau)
- ( S^*C + SC^*)\cos(2\omega\,\tau)\Big]\nonumber\\
&+ \frac{1}{\omega^2} \left(|S|^2 - |C|^2\right) + |S'|^2 + |C'|^2 \,,
\label{eq:E}
\end{align}
where the asterisk denotes complex conjugation.
Using these expressions, it is easy to see that the energy balance
equation $d_t E = q \Omega_0 T_{xy}$ is also satisfied up to order $1/\omega\,\tau$.
In Figure~\ref{fig:se}, we illustrate the numerical solutions for the stress
$T_{xy}(t)$ and energy $E(t)$, given by the thin and thick solid
lines, together with the analytical approximation for the energy.
The latter has been obtained by substituting the two pairs of
constants, $C_- = 1$ and $S_- = 0$, and $C_+ = 0.285$ and $S_+ =
-1.896$, into equation~(\ref{eq:E}).
It is thus clear that the late-time stress oscillates around zero with
decreasing amplitude, while the energy density asymptotes
to a non-vanishing, time-independent value.
The expression for the energy density at early/late times in terms of
the constants $S_\pm$ and $C_\pm$ is given by\footnote{Here, we assume
$S' = C' = 0$, and thus avoid the uninteresting contributions due to
any stable Alfv\'en wave initially present, see
equations~(\ref{sol:uz})--(\ref{sol:bz}).}
\begin{align}
E_\pm \equiv \lim_{t\rightarrow\pm\infty} E(t) =
\frac{|S_\pm|^2 + |C_\pm|^2}{\omega^2}\,.
\label{eq:amplification}
\end{align}
Therefore, the energy gain via swing amplification, $E_+/E_-$, is in
general a function of the ratio $\omega = \omega_\mathrm{A}/q\,\Omega_0$ and the
initial conditions.
However, it is possible to obtain conclusions that are independent of
the latter.
The dependence of the energy gain on the
initial conditions for $\omega^2= 1$ is shown in Figure~\ref{fig:rs}.
The horizontal axis describes how the initial energy is distributed
between the $j_n$ modes and the $y_n$ modes; while the different lines
show the phase difference in the corresponding initial amplitudes.
When $\arg(S_-/C_-) = \pi/2$ or $3\pi/2$, the $j_n$ and $y_n$ modes
are completely out of phase and evolve independently.
This results in an energy gain which is linear in the initial
amplitudes (thick solid line).
We have found that the dependence of the energy gain on the phase
difference between the constants determining the initial conditions is
weaker if $\omega$ decreases below unity.
In this case, all the different curves converge to the thick line
corresponding to $\arg(S_-/C_-) = \pi/2$.
At the same time, as $\omega$ decreases below unity, this line gets
steeper, providing thus a larger energy gain.
This justifies referring to the $y_n$ and $j_n$ as the ``growing'' and
``decaying'' modes, respectively.
We illustrate the dependence of the energy gain on the shear parameter in Figure~\ref{fig:rq}
(because the results depend only on $\omega^2$, we only show the positive domain
in the horizontal axis). In the limit of weak shear, there are only pure Alfv\'en waves and
there is no net energy gain.
The dashed line shows that the energy gain tends to the value $E_+/E_-
= 10/\omega^2$ as $1/\omega \gg 1$, which provides a good description
of the numerical results for strong shear.
The asymptotic behavior is insensitive to the initial conditions as
long as the growing mode is excited, i.e. $C_-\ne 0$.
\begin{figure}
\includegraphics[width=\columnwidth,trim=16 8 8 16]{f3.eps}
\caption{The solid curves show the dependence of the energy gain,
$E_+/E_-$, on the initial conditions for $\omega = \omega_\mathrm{A}/q\,\Omega_0
= 1$.
The horizontal axis provides a measure of the relative amplitude
of $C_-$ and $S_-$; while the different lines show the results for
various phase differences.
For strong shear $1/\omega^2 > 1$, the dependence of the phase is
weaker and all the different curves collapse onto the straight thick
line, which corresponds to $\arg(S_-/C_-) = \pi/2$, while this one
gets steeper, thus providing a larger energy gain, see also
Figure~\ref{fig:rq}.}
\label{fig:rs}
\end{figure}
\section{Discussion}
\subsection{Summary and Connection to Previous Work}
We have employed the shearing-sheet framework to study the dynamical
evolution of MHD waves in weakly magnetized differentially rotating
backgrounds which are stable to the MRI. While the fact that these
waves can be transiently amplified is widely appreciated,
our motivation to study them, as well as the
results that we obtained, concern dynamical aspects that have not
received as much attention. This is whether these shearing MHD waves
can play a significant role in the transport of angular momentum in regions of
the disk where the MRI is inefficient, such as the accretion disk boundary layer.
\begin{figure}
\includegraphics[width=\columnwidth,trim=16 8 8 16]{f4.eps}
\caption{The filled circles represent the value of the energy gain $E_+/E_-$ for
different values of the shear, parameterized via $q\Omega_0/\omega_\mathrm{A} =
1/\omega$, obtained via numerical integration using the initial
conditions $C_- = 1$ and $S_- = 0$, i.e., only the growing mode is
excited.
In the limit of weak shear, there are only pure Alfv\'en waves and thus
there is no net energy gain.
The dashed line shows the function $10/\omega^2 \equiv
10(q\Omega_0/\omega_\mathrm{A})^2$, which is in good agreement with the
numerical results for strong shear.
This asymptotic behavior is independent of the initial conditions
as long as $C_- \ne 0$.}
\label{fig:rq}
\end{figure}
The equations that we have solved are similar to those presented in
\citet{1992ApJ...400..610B}, who provided numerical solutions showing
that transient amplification of MHD waves is a general outcome for the
modes with wavevectors that are not exactly aligned with the rotation
$z$-axis\footnote{For an analysis of non-axisymmetric spiral waves
when only a strong, vertical background magnetic field is considered
see \citet{1992ApJ...393..708T}.}.
Analytical insight on the non-linear dynamics of these waves has been
usually hindered by the fact that the governing equations cannot be
simplified beyond a set of coupled differential equations (see, e.g.,
\citealt{1997MNRAS.291...91F, 2000ApJ...540..372K,
2006A&A...450..437B}, and also \citealt{2007ApJ...660.1375J}, where
higher order WKB solutions for the linear evolution of the shearing
waves are provided).
In this paper, by isolating the modes that are unrelated to the
standard MRI, i.e., by setting $k_z=0$, we have been able to provide
analytical solutions that are valid for all times, except close to the
instant where the waves evolve from leading to trailing, i.e., when
$k_x(t)=0$.
These solutions are exact when only one Fourier mode is considered
\citep{1994ApJ...432..213G}.
Despite the fact that we do not predict analytically the amplification
factor for these waves, the characteristics of the solutions that we
found allowed us to draw important conclusions.
We showed that the amplification factor is \emph{only} a function of
the (time-independent) dimensionless ratio $(q\Omega_0/\omega_{\rm
A})^2$, with
\begin{align}
\frac{E_+}{E_-} \approx 10 \left(\frac{q\Omega_0}{\omega_{\rm A}}\right)^2 \quad
\textrm{for} \quad q\Omega_0 \gg \omega_{\rm A} \,,
\end{align}
see equation~(\ref{eq:amplification}) and Figure~\ref{fig:rq}, and it is
thus insensitive to the sign of the shear parameter $q$.
An important result of this study is that while the energy of these
MHD waves can be significantly amplified, their net associated stresses
oscillate around zero, see equation~(\ref{eq:Txy}) and Figure~\ref{fig:se}.
This suggests that these shearing MHD waves are unlikely to play an important
role in the transport of angular momentum in the accretion disk boundary layer region.
These findings are consistent with the results of global MHD simulations
of accretion disks with a rigid inner boundary carried out in
\citet{2002MNRAS.330..895A} and \citet{2002ApJ...571..413S}.
These simulations show that the inner disk regions, where $d\Omega/dr \ge 0$,
can develop strong toroidal magnetic fields, with associated magnetic energies
that can easily reach a few tenths of the thermal energy, without leading to
efficient angular momentum transport.
\subsection{Implications}
The importance of understanding the relationship between the stress
and the radial gradient in angular frequency resides in that this
dependence plays a key role when modeling the inner structure of an
accretion disk surrounding a weekly magnetized star (see, e.g.,
\citealt{1995ApJ...442..337P, 1996ApJ...473..422P}).
In the steady state, the constant inward flux of angular momentum at
any given radius $r_0$ is given by $\dot J = \dot M l - 2\pi r_0^2 H
T_{r\phi}$, where $\dot M$ stands for the accretion rate, $l$ is the
specific angular momentum, $H$ is the disk height, and $T_{r\phi}$
(denoted by $T_{xy}$ in our analysis) is the component of the stress responsible for the
flux of azimuthal momentum across the radial direction.
Thus, the angular momentum flux has two contributions: $\dot
J_\mathrm{matter} = \dot M l$, which accounts for the flux of angular
momentum due to mass accretion, and $\dot J_\mathrm{stress} = - 2\pi
r_0^2 H T_{r\phi}$, which accounts for the flux of angular momentum
due to the stress acting on the fluid elements constituting the disk.
Under the reasonable assumption that the disk must be Keplerian
well beyond the boundary layer, i.e.,
for $r \gg r_\mathrm{b}$, and that the stress should vanish in the absence of shear, we
must have $\dot J \equiv \dot M l(r_\mathrm{b})> 0$, i.e., a slowly rotating
star accretes mass {\it and} angular momentum.
In order for this picture to be self-consistent, the stress must
satisfy $T_{r\phi}(r_\star \le r\le r_\mathrm{b}) \le 0$.
In the standard accretion disk model this requirement is satisfied by
assuming that the stress is linearly proportional to the local shear
$T_{r\phi} \sim - d\Omega/dr$.
With this model for the stress, and some supplementary assumptions, it
is possible to solve for the radial dependence of $\Omega(r)$ and
determine the structure of the disk (see, e.g.,
\citealt{1991ApJ...370..604P}).
However, this assumption, broadly adopted in the framework of enhanced
turbulent disk viscosity, does not seem to be supported by the modern
paradigm, in which angular momentum transport is due to MHD turbulence
driven by the MRI. Indeed, both local numerical simulations of
shearing-boxes with non-Keplerian shear profiles
\citep{2008MNRAS.383..683P, 2009A&A...505..955S} and
global disk simulations with a rigid inner boundary
\citep{2002MNRAS.330..895A, 2002ApJ...571..413S} suggest that angular
momentum transport is inefficient if $d\Omega/dr > 0$.
This suggests that the detailed structure of accretion disk boundary layers
resulting from the interaction of an MHD disk with a weakly magnetized star
could differ appreciably from those derived within the standard turbulent shear
viscosity, where the direction of angular momentum transport
is always opposite to the angular frequency gradient.
\subsection{Final Remarks}
It is worth mentioning explicitly that the shearing-sheet framework that
we have employed is inherently limited to address the conditions
expected in the accretion disk boundary layer. For example, the
absence of a hard-inner boundary could prevent Kelvin-Helmholtz
instabilities from operating, see, e.g., \citet{2009ApJ...702.1536B}.
However, these instabilities do not seem to play a predominant role in
the global MHD simulations of \cite{2002MNRAS.330..895A} and
\cite{2002ApJ...571..413S}. Moreover, because we have assumed a constant
background density, our analysis precludes the possibility
of buoyant modes or convective instabilities. Whether these instabilities,
and the turbulence they could drive, transport angular
momentum inward or outward in Keplerian disks has been long debated
\citep{1992ApJ...388..438R, 1996ApJ...465..874C, 1996ApJ...464..364S, 2010MNRAS.404L..64L}.
To our knowledge, these convective instabilities have
not been studied in differentially rotating backgrounds with angular frequencies
increasing outward; and speculating about their role goes beyond the scope of
the present work.
In spite of the simplifications of our analytical study,
the explicit solutions that we have found can provide physical insight and
help elucidate transport processes in the inner disk regions close to a weakly
magnetized accreting star.
The current availability of powerful parallel codes (e.g. \citealt{2008ApJS..178..137S})
which are already being used to study the hydrodynamics of accretion disk boundary
layers \citep{2011arXiv1112.3102B} holds the promise
that a more detailed understanding of MHD boundary layers will soon
be possible.
\acknowledgments
MEP is grateful to the Knud H{\o}jgaard Foundation for its generous support.
CKC is supported by a NORDITA fellowship.
We thank Tobias Heinemann, John Wettlaufer, and Jim Stone for useful discussions.
|
\section{Introduction}
Random matrix models whose eigenstates exhibit a transition from extended to localized states provide an efficient tool for studying the Anderson metal-insulator transition \cite{Mut93, MNS94, MFD96, KM97, FOR09, PPTW11, EM08}. Their main advantage, compared to the original Anderson model \cite{And58}, is that the transition occurs not just at a single point in the parameter space, but rather on a critical line described by the variation of an additional parameter, such as the band width of the power-law banded random matrix model \cite{MFD96}. The models are accessible to perturbative treatment when this parameter is either large or small \cite{ME00, CK07, KOY11, ROF11, MG10}.
So far, most of the critical random matrix models which have been studied intensively are one-dimensional. Their Hamiltonians describe random hopping of a particle on a one-dimensional lattice in a random on-site potential.
The existence of the Anderson transition in such one-dimensional systems is related to the long-range nature of the hopping amplitudes. In our recent work \cite{ORC11}, we studied the scaling of the moments of the eigenstates in a {\it two-dimensional} generalization of the power-law banded random matrix model \cite{PS02,C04}. In this ensemble, the matrix elements of the Hamiltonian $H_{{\bf m}{\bf n}}$ are complex independent Gaussian random variables, whose mean values are equal to zero and whose variances are determined by the distance between sites of a two-dimensional lattice:
\begin{equation}\label{Ham}
\left\langle|H_{{\bf m}{\bf n}}|^2\right\rangle\equiv \frac{1}{1+(|{\bf m}-{\bf n}|/b)^4},
\end{equation}
where ${\bf m}=(m_x,m_y),\;{\bf n}=(n_x,n_y),\;1\le m_{\alpha},n_{\alpha}\le L$ are two-dimensional vectors representing two sites on a two-dimensional square lattice of size $L\times L$, and $b$ is a parameter of the model. Thus the Hamiltonian is represented by a random $L^2\times L^2$ matrix.
One natural way to convince oneself that this system is critical is to study how the moments of its eigenfunctions $\psi_n({\bf r})$ scale with the system size $L$. Namely, we define
\begin{equation}\label{mom_def}
I_q=L^{d}\left\langle|\psi_n({\bf r})|^{2q}\right\rangle\propto L^{-d_q(q-1)},
\end{equation}
where $d$ is the dimensionality of the space and the averaging is performed over the ensemble as well as over a small energy window. Trivial exponent values $d_q=0$ and $d_q=d$ signify localized and extended states, respectively. For critical states, $0<d_q<d$ and what is more, $d_q$ generally depends on $q$, indicating that the eigenfunctions are multifractal. The above scaling of the moments of the eigenfunctions was obtained analytically and confirmed by numerical simulations in various critical random matrix ensembles \cite{EM08}. In particular, it was found that in the power-law banded random matrix model $d_q\ll 1$ when the bandwidth $b\ll 1$ (strong multifractality) and $d-d_q\ll1$ when $b\gg 1$ (weak multifractality) \cite{MFD96,ME00}.
One of the surprising features of the model (\ref{Ham}) revealed in \cite{ORC11} was the absence of a pure power-law scaling of the moments of the eigenfunctions at $b\gg 1$. Instead of Eq.(\ref{mom_def}), the scaling was surmised to be
\begin{equation}\label{ln_scaling}
I_q\propto L^{-2(q-1)}\ln^{\nu_q(q-1)} L,
\end{equation}
where the exponents $\nu_q$ play the role of the anomalous fractal dimensions $d-d_q$ and can be calculated perturbatively in a similar way \cite{ORC11}. In the same time, the standard power-law scaling (\ref{mom_def}) with $d_q\ll 1$ was found at $b\ll 1$, in full analogy with the one-dimensional version of the model.
The existence of two different scaling laws at large and small values of $b$ suggests that there should be a critical value of $b=b_{\rm cr}$ separating these two regimes. One of the aims of the present paper is to show that such a critical value of $b$ does exist. While the model can be treated analytically at $b\ll 1$ or $b\gg 1$, the existence of the transition between two regimes can be investigated only with the help of numerical simulations. These are more efficient for spectral properties rather than for the statistics of the eigenvectors. For this reason, in the present work we focus on studying the spectral compressibility.
The spectral compressibility $\chi$ is defined by the asymptotic behavior of the level number variance:
\begin{equation}\label{comp-def}
\langle \delta n^2(E)\rangle=\langle n^2(E)\rangle -\langle n(E)\rangle^2\approx \chi \langle n(E)\rangle,\quad \langle n(E)\rangle\gg 1,
\end{equation}
where $n(E)$ is the number of eigenvalues in a spectral window of the width $E$. It is well known that $\chi$ plays the role of a critical exponent: $\chi=0$ in the metallic phase, $\chi=1$ in the localized phase and $0<\chi<1$ at criticality \cite{EM08}.
Our main results concerning the behavior of the level number variance and the compressibility in the two-dimensional model (\ref{Ham}) can be formulated as follows. The asymptotic behavior of $\langle \delta n^2(E)\rangle$ is described by Eq.(\ref{comp-def}) for all values of $b$ such that $0\le b\le b_{\rm cr}$. The spectral compressibility $\chi(b)$ is a monotonically decaying function of $b$ with $\chi(0)=1$ and $\chi(b_{\rm cr})=0$. The latter equation is used as the definition of $b_{\rm cr}$. At small values of $b$, $\chi(b)$ can be calculated perturbatively:
\begin{equation}\label{comp_small_b}
\chi(b)=1-\frac{\pi^2 b^2}{\sqrt{2}}+O(b^4).
\end{equation}
\begin{figure}[t]
\begin{center}
\includegraphics[clip=true,width=\columnwidth]{var-num10.eps}
\end{center}
\caption{Level number variance as a function of the average level number, for different values of $b<1$ and $L=64$.}
\label{Fig1}
\end{figure}
For $b\ge b_{\rm cr}$, spectral compressibility is equal to zero. The level number variance in this case has the same asymptotic behavior as in the Wigner-Dyson random matrix theory:
\begin{equation}
\langle \delta n^2(E)\rangle=\frac{1}{\pi^2}\left(\ln (2\pi \langle n\rangle)+\gamma+1\right),\quad \langle n\rangle\gg 1,
\end{equation}
where $\gamma$ is the Euler constant. Our analysis, based on numerical simulations, shows that the transition between these two regimes occurs at $b_{\rm cr}=5.2 \pm 0.2$.
The remainder of this paper has the following structure. In Section \ref{sec_small_b} we present a derivation of Eq.(\ref{comp_small_b}) at $b\ll 1$ and confirm the obtained result by numerical simulations. The opposite case of $b\gg 1$ and the transition between the two phases are considered in Section \ref{sec_large_b}. We summarize our results in Section \ref{conclusions}. Finally, the Appendix contains an alternative derivation of Eq.(\ref{comp_small_b}), easily to extended to the orthogonal symmetry class.
\section{Spectral compressibility at $b\ll 1$}\label{sec_small_b}
A typical random matrix from the ensemble (\ref{Ham}) has the diagonal elements of order one and the off-diagonal elements of order $b$. Thus, the off-diagonal elements are parametrically smaller than the diagonal ones provided that $b\ll 1$. This allows developing a perturbation expansion of the eigenfunction moments, spectral compressibility or any other quantity of interest in a power series of $b$. In Ref.\cite{ME00,YK03} this method was applied to the calculation of spectral compressibility in the one-dimensional power-law banded random matrix model. Below we use the same approach in the two-dimensional model (\ref{Ham}).
\begin{figure}[t]
\begin{center}
\includegraphics[clip=true,width=\columnwidth]{chivsb2-12.eps}
\end{center}
\caption{Spectral compressibility as a function of $b$, obtained from the slopes of the lines in Fig.~\ref{Fig1}. The data is presented for three different system sizes. The error bars are smaller than the symbol sizes.}
\label{Fig2}
\end{figure}
The zeroth order term in the expansion corresponds to a pure diagonal random matrix, which has completely localized eigenvectors and uncorrelated eigenvalues. As a result, $\chi=1$ in this case. The first order correction to this trivial result can be found from the relation between $\chi$ and the form factor $K^{(1)}(t,N)$\cite{YK03}:
\begin{equation}
\chi=1+\lim_{t\to\infty}\lim_{N\to\infty}K^{(1)}(t,N),
\end{equation}
where $N$ is the matrix size, which is equal to $L^2$ in our case, and the form factor is calculated in the lowest order of the perturbation theory. The general expression for the latter reads \cite{YK03, YO07,footnote}
\begin{eqnarray}
\label{sum_m}K^{(1)}(t,N)&=&-\frac{2\sqrt{\pi}}{t}\sum_{{\bf m}\neq 0}x(|{\bf m}|)e^{-x(|{\bf m}|)},\\
x(|{\bf m}-{\bf n}|)&=&\frac{1}{2}\langle|H_{{\bf m}{\bf n}}|^2\rangle t^2.
\end{eqnarray}
Since we are interested in the limit $N\to\infty$, the sum over ${\bf m}$ in Eq.(\ref{sum_m}) can be replaced by a two-dimensional integral, which can be transformed to the polar coordinates upon substitution of $\langle|H_{{\bf m}{\bf n}}|^2\rangle$ given in (\ref{Ham}):
\begin{equation}
\lim_{N\to\infty}K^{(1)}(t,N)=-2\pi^{3/2}t\int_{1}^{\infty}dr \frac{r e^{-\frac{t^2}{1+(r/b)^4}}}{1+(r/b)^4}.
\end{equation}
This integral can be evaluated in the limit $t\to\infty$ by changing the variable from $r$ to $s=b^4t^2/2r^4$:
\begin{equation}\label{correction}
\lim_{t\to\infty}\lim_{N\to\infty}K^{(1)}(t,N)=-\frac{\pi^{3/2}b^2}{\sqrt{2}}\int_{0}^{\infty}\frac{e^{-s}}{\sqrt{s}}= -\frac{\pi^{2}b^2}{\sqrt{2}},
\end{equation}
leading to the formula announced in Eq.(\ref{comp_small_b}). The same result can be obtained directly by calculating $\chi$ for a random $2\times2$ matrix, as shown in the Appendix. The corresponding expression for $\chi$ in the orthogonal symmetry class (real symmetric matrices) is also given there.
In order to test this prediction we performed numerical simulations with random matrices generated according to Eq.(\ref{Ham}). The spectra of these matrices were obtained by standard diagonalization subroutines \cite{LAPACK}, and the number of realizations for each $b$ was $20000$, $2000$ and $500$ for $L = 64$, $128$ and $180$, respectively.
The numerical findings for $b<1$ are summarized in Fig.~\ref{Fig1} and Fig.~\ref{Fig2}. The first one shows the level number variance $\langle \delta n^2(E)\rangle$ as a function of $\langle n(E)\rangle$ for different values of $b$. The linear behavior suggested by Eq.(\ref{comp-def}) is evident for $b<1$. Extracting the slopes of the straight lines we obtain the numerical values for $\chi(b)$, which are presented in Fig.~\ref{Fig2}. They are in a good agreement with our analytical prediction (\ref{comp_small_b}), which is valid only when the absolute value of the perturbative correction (\ref{correction}) is much smaller than one, i.e. $b^2\ll \sqrt{2}/\pi^2\approx 0.14$.
\section{Level number variance at $b\gg 1$ and the transition point}\label{sec_large_b}
To study the limit $b\gg 1$ we exploit the fact that in this case the model can be mapped onto a non-linear $\sigma$-model \cite{ORC11}. In the $\sigma$-model description the key ingredient characterizing a particular model is the propagator. In Ref.\cite{ORC11} it was found that the propagator of the two-dimensional model (\ref{Ham}) is given by
\begin{equation}\label{prop}
\Pi(k)=-\frac{\pi^3}{2}\frac{1}{k^2\ln bk}
\end{equation}
in the momentum space. Its eigenvalues determine various spectral properties of the model \cite{KL95,M00}. In particular, they allow us to calculate the two-level correlation function
\begin{equation}\label{R}
R(\omega)=\frac{\Delta^2}{2\pi^2}{\rm Re}\,\sum_{{\bf k}}\frac{1}{\frac{2 b^2}{\pi^2 \nu}k^2\ln bk-i\omega},
\end{equation}
where $\Delta$ is the mean level spacing, $\nu=1/\Delta L^2$ is the density of states and the sum runs over the discrete momenta ${\bf k}=(2\pi m_x/L,2\pi m_y/L),\: m_{\alpha}\in \mathbb{Z}$. The two-level correlation function allows us to find the derivative of $\langle \delta n^2(E)\rangle$ \cite{KL95}
\begin{equation}\label{common}
\frac{d\langle \delta n^2\rangle}{d\langle n\rangle}=\lim_{L\to \infty}\int_{-\langle n\rangle}^{\langle n\rangle}ds\: R(s), \quad s=\omega/\Delta.
\end{equation}
Substituting the expression (\ref{R}) into this formula and integrating over $s$ we obtain
\begin{equation}
\frac{d\langle \delta n^2\rangle}{d\langle n\rangle}=\frac{\langle n\rangle}{(8\pi b^2)^2}\lim_{L\to \infty}\sum_{{\bf m}}\frac{1}{m^4\ln^2\left(\frac{2\pi b m}{L}\right)+\left(\frac{\langle n\rangle}{8b^2}\right)^2}.
\end{equation}
Since the sum over ${\bf m}=(m_x,m_y),\: m_{\alpha}\in \mathbb{Z}$ converges and each term with ${\bf m}\neq (0,0)$ tends to zero in the limit $L\to \infty$, we conclude that only the zero mode ${\bf k}=0$ has a non-vanishing contribution in the thermodynamic limit.
It is well known that the contribution of the zero mode of the $\sigma$-model reproduces the results of the Wigner-Dyson random matrix theory \cite{Efetov}. Thus we must expect that in this regime $\langle \delta n^2(E)\rangle$ is given by the standard random matrix theory result \cite{Mehta}:
\begin{equation}\label{rmt-res}
\langle \delta n^2(E)\rangle=\frac{1}{\pi^2}\left(\ln (2\pi \langle n\rangle)+\gamma+1\right),\quad \langle n\rangle\gg 1.
\end{equation}
The absence of the linear in $\langle n\rangle$ term in this asymptotic law implies that $\chi=0$, as one usually finds in the metallic phase. We would like to point out that similar calculations for the one-dimensional version of the model give a non-zero result for $\chi$ \cite{MFD96, KM97}. It is the logarithmic term in the propagator (\ref{prop}), that leads to the vanishing of $\chi$. The same logarithmic term is responsible for the unusual scaling behavior of the moments of the eigenfunctions (\ref{ln_scaling}).
\begin{figure}[t]
\begin{center}
\includegraphics[clip=true,width=\columnwidth]{var-num1.eps}
\end{center}
\caption{Level number variance as a function of the average level number, for different values of $b>1$ and $L=64$. Dots are numerical data. The solid line at $b=5$ is Eq. (\ref{rmt-res}). The other solid lines are the results of fitting the data with Eq. (\ref{interpol}).}
\label{Fig3}
\end{figure}
Fig.~\ref{Fig3} shows the results of numerical simulations for $b\ge 1$. We observe a very good agreement between Eq. (\ref{rmt-res}) and the numeric data at $b=5$. One can also see how the behavior of $\langle \delta n^2\rangle$ changes gradually from linear for small $b$ to logarithmic for large $b$. To quantify this change, as well as to determine the transition point between the two regimes, we assume that the most general functional form of $\langle \delta n^2\rangle$ is given by
\begin{equation}\label{interpol}
\langle \delta n^2\rangle=\chi\langle n\rangle+\frac{a_1}{\pi^2}\ln(2\pi\langle n\rangle)+a_2, \quad \langle n\rangle\gg 1.
\end{equation}
This formula naturally interpolates between Eq. (\ref{comp-def}) and Eq. (\ref{rmt-res}). Moreover we note that Eq.(\ref{comp-def}) is never realized in its pure form: there is always a subleading logarithmic term. If the transition between the two regimes occurs at some finite value of $b=b_{\rm cr}$, one should expect that $\chi$ decays as a function of $b$ and goes to zero at $b_{\rm cr}$.
Using $\chi$, $a_1$ and $a_2$ as fitting parameters, we were able to reproduce the behavior of $\langle \delta n^2\rangle$ at intermediate values of $b$ as shown in Fig.~\ref{Fig3}. Moreover, the best-fit parameters for large values of $b$ are in agreement with Eq.(\ref{rmt-res}). At $b=5$, for example, we find $\chi= 0.00012 \pm 0.00016$, $a_1= 1.007 \pm 0.007$ and $a_2= 0.1573 \pm 0.003$, while the standard random matrix theory prediction (\ref{rmt-res}) corresponds to $\chi=0$, $a_1=1$ and $a_2=0.1598$.
The values of $\chi$, obtained in this way for different values of $b$, are presented in Fig.~\ref{Fig4}. One can see that, indeed, $\chi$ monotonically decreases as a function of $b$ and becomes zero within numerical accuracy for $b\ge b_{\rm cr}$. A close study of $\chi$ in the interval $4.5\le b\le 5.5$ allowed us to determine the critical value $b_{\rm cr}= 5.2 \pm 0.2$.
\section{Conclusions}\label{conclusions}
The random matrix model described by Eq.(\ref{Ham}) was studied numerically in Ref.\cite{PS02, C04}. The results of those works suggest that the model is critical at all values of $b$. All analytical and numerical findings known for the one-dimensional counterpart of the model support the same expectation \cite{MFD96, EM08}.
In this paper we show that the situation is actually more subtle. For $b \ll 1$ we found the expected critical behavior, characterized by a non-zero value of the spectral compressibility (\ref{comp_small_b}). Also, the scaling of the eigenstates is the standard power-law (\ref{comp-def}) with non-trivial multifractal dimensions $d_q$. For $b \gg 1$, on the other hand, we obtained $\chi=0$. This is normally a signature of the standard metallic phase corresponding to completely extended states with $d_q=d$. However, this is not the case in our model, where the moments of the eigenstates contain an additional anomalous part which scales as a power of the logarithm of system size (\ref{ln_scaling}). Therefore, the phase at $b\gg 1$ is not entirely metallic. We may say that it has some traces of the critical behavior. We found that the transition between the two phases -- ``critical" and ``metallic-critical" -- occurs at $b_{\rm cr}= 5.2\pm 0.2$.
We believe that two important features of the model are responsible for the emergence of the two phases: the dimensionality $d=2$ and the long-range nature of the Hamiltonian (\ref{Ham}). Recently, a similar behavior was predicted for another two-dimensional long-range random Hamiltonian \cite{AAE11}.
\begin{figure}[t]
\begin{center}
\includegraphics[clip=true,width=\columnwidth]{a0vsb-10.eps}
\end{center}
\caption{Spectral compressibility as a function of $b$, extracted from the data in Fig.~\ref{Fig3} using Eq. (\ref{interpol}). The error bars are smaller than the symbol size.}
\label{Fig4}
\end{figure}
Finally, we would like to point out that $\chi$ given in Eq.(\ref{comp_small_b}) and $d_1=\sqrt{2}\pi^2b^2$, as calculated in Ref.\cite{ORC11}, satisfy the relation
\begin{equation}
\chi+\frac{d_1}{d}=1,
\end{equation}
which was suggested recently in Ref.\cite{BG11} and verified for various one-dimensional random matrix models.
\section{Acknowledgments}
We are grateful to Yan Fyodorov, Vladimir Kravtsov and Alexander Mirlin for useful discussions. IR and AO acknowledge support from the Engineering and Physical Sciences Research Council [grant number EP/G055769/1]. E.C. thanks the FEDER and the Spanish DGI for financial support through Project No. FIS2010-16430.
\section{Appendix. Calculation of spectral compressibility at $b\ll1$.}\label{app}
In the limit of an almost diagonal Hamiltonian $H_{ij}$ the problem can be reduced to a $2\times 2$ Hamiltonian \cite{EM08}:
\begin{equation} H_2=\left(
\begin{array}{cc}
\epsilon_1 & b h \\
b h^* & \epsilon_2 \\
\end{array}
\right),
\end{equation}
whose eigenvalues $\lambda_{\pm}$ are the solutions of the characteristic equation
\begin{equation}\label{characteristic} (\epsilon_1-\lambda)(\epsilon_2-\lambda)-b^2|h|^2=0.\end{equation} Both $\epsilon_{1,2}$ are Gaussian with variance $\sigma^2$.
The density-density correlator of this $2\times 2$ matrix correlator near $E=0$ is
$R(\omega)=\frac{1}{4}\left\langle \delta(\omega-\lambda_+)\delta(\lambda_-)\right\rangle.$ We are interested in the true density correlator of the original large random matrix. Although the correlators in these two cases are related, they are not the same: it turns out that we need to double the $2\times2$ $R(\omega)$. Indeed, in a very large matrix diagonal matrix (the limit $b=0$) $\langle\rho(\omega)\rho(0)\rangle=\langle
\rho(\omega)\rangle\langle\rho(0)\rangle$, whereas in the $2\times 2$ matrix the r.h.s. contains a coefficient $1/2$. Doubling the correlator, we have
\begin{equation} R(\omega)=\frac{1}{8\pi \sigma^2}\left\langle\int e^{-\frac{\epsilon_1^2+\epsilon_2^2}{2\sigma^2}}\delta(\omega -\lambda_+)\delta(\lambda_-)d\epsilon_1 d\epsilon_2 \right\rangle_h,\end{equation}
where $\langle\rangle_h$ is the average over the off-diagonal elements. Using the Eq.. (\ref{characteristic}) we can change the integration variable $\epsilon_1\to \lambda_-$. One integral is then removed by $\delta(\lambda_-).$ The other $\delta$-function ensures $\epsilon_2+\frac{b^2|h|^2}{\epsilon_2}=\omega$, which can be used in the exponential. Afterwards it becomes clear that the exponential can be neglected, as long as we are working in an energy window $\omega\ll\sigma$. In this limit,
\begin{equation} R(\omega)=\frac{1}{4\pi \sigma^2}\left\langle \frac{|\omega|}{ \sqrt{\omega^2-4b^2|h|^2}}\mathbf{1}_{|\omega|>2b|h|}
\right\rangle_h.\end{equation} The average density of states at $E=0$ is given by a simpler and similar calculation:
$\langle\rho(0)\rangle=\frac{1}{\sqrt{2\pi \sigma^2}}.$
Knowing the density correlator in the $2\times 2$ case, we can find the leading order of the density correlator in the original large system by replacing $bh$ with the off-diagonal elements $H_{ij}$ and introducing the summation over $i,j$. Then we can substitute the connected correlator into Eq. (\ref{common}). As a result,
\begin{equation}\label{the amazing formula} \chi=1-\lim_{\langle n\rangle\to\infty}\Bigl(2\langle n\rangle-\nonumber\end{equation}
\begin{equation}\frac{2\langle n\rangle}{N^2}\sum_{i\neq j}\Bigl\langle\sqrt{1- \frac{2N^2}{\pi \langle n\rangle^2}\frac{1}{\sigma^2}|H_{ij}|^2} \mathbf{1}_{|H_{ij}|^2<\pi\sigma^2 \frac{\langle n\rangle^2}{2N^2}}\Bigr\rangle\Bigr).\end{equation}
This result is general -- valid for any dimensionality and for any symmetry class. Note that each term of the sum is an independent random variable and so can be averaged separately. Performing the averaging in the unitary case, we obtain $\chi=1-\lim_{\langle n\rangle\to\infty}I$ where
\begin{equation} I^{(U)}=\sum_{\bf{r},\bf{r}'}\frac{\sqrt2a_{{\bf r}, {\bf r}'}}{N} \exp\!\left(-\frac{\pi \langle n\rangle^2}{2a_{{\bf r}, {\bf r}'}^2N^2}\right)\!\,{\mathrm{erfi}}\!\left(\sqrt\frac{\pi}{2}\frac{\langle n\rangle}{a_{{\bf r}, {\bf r}'}N}\right).\end{equation}
Here we used the standard notation for the variance $\langle|H_{{\bf r}, {\bf r}'}|^2\rangle=a^2_{{\bf r}, {\bf r}'}$.
In our 2D system it is sufficient to set $a_{{\bf r}, {\bf r}'}=\frac{b^2}{|{\bf r}- {\bf r}'|^2}$. To simplify the calculation, we can imagine that our system occupies a very large disk of radius $R$ (so that $N=\pi R^2$), replace summation with integration over ${\bf r}$ and ${\bf r}'$ and subsequently change the integration variables to ${\bf r}- {\bf r}'$ and ${\bf r}+{\bf r}'$. The integrals can be computed in polar coordinates. The first integration depends weakly on the large upper limit and so we can set it to $R$. Afterwards the integrand does not depend on ${\bf r}+{\bf r}'$, making the second integration trivial. The result is
\begin{equation} I^{(U)}\approx 2\langle n\rangle\,_pF_q\left(\frac{1}{2},1;\frac{3}{2}, \frac{3}{2};-\frac{\langle n\rangle^2}{2\pi b^4}\right).\end{equation}
Taking the large $\langle n\rangle$ limit,
\begin{equation}\chi^{(U)}=1-\frac{\pi^2 b^2}{\sqrt2}+O(b^4).\end{equation}
The case of the orthogonal symmetry is treated in precisely the same way, starting from Eq. (\ref{the amazing formula}). There we obtain
\begin{equation}\chi^{(O)}=1-2\pi b^2+O(b^4).\end{equation}
|
\section{Introduction}
\label{sec:intro}
The matter that is directly observed in nature
consists of atoms, whose nuclei are droplets of nuclear matter
composed of up and down quarks. Nuclear matter is very stable:
the most stable nuclei have lifetimes longer than the age of the universe.
However, it has been hypothesized
\cite{Bodmer:1971we,Witten:1984rs,Farhi:1984qu}
that nuclear matter may actually be metastable, and the
true ground state of matter consists of a combination of roughly
equal numbers of up, down, and strange quarks known as ``strange matter''.
Strange matter is hypothesized to exist as
(kilometer-sized) pieces, known as ``strange stars'' (reviewed in
\cite{Weber:2004kj}), or as small nuggets, known as ``strangelets''
\cite{Farhi:1984qu}.
It has further been hypothesized that dark matter could be some
form of quark matter, trapped in strangelets or strange stars before
the era of nucleosynthesis
\cite{Zhitnitsky:2002qa,Banerjee:2002et,Forbes:2009wg}.
However, even if strange matter is not invoked as a dark matter candidate,
there could still be a population of strange matter objects, from strange
stars to strangelets, many of which
would be relatively non-luminous. In this article we
show that the masses and radii of such objects can extend in to the range
expected for planets. Recent surveys such as the Microlensing Observations in Astrophysics (MOA) and
the Optical Gravitational Lensing Experiment (OGLE) to detect such low mass non-luminious low mass objects
by gravitational lensing have yielded interesting results \cite{Sumi:2011}.
The hypothetical compact objects we predict
could be detected by such methods and such surveys could place stringent bounds or perhaps hint at their
possible existence.
It is generally assumed that strange stars are compact objects, with sizes in
the 10 kilometer range \cite{Weber:2004kj}, ending at a sharp surface of
thickness $\sim$ 1 Fermi, perhaps with a very thin electrostatically suspended
nuclear matter crust~\cite{Alcock:1986hz,Stejner:2005mw,Usov:1997eg}.
However, if the surface tension $\sigma$ of the interface between quark matter
and the vacuum is less than a critical value $\si_{\rm crit}$
(of order a few MeV/fm$^2$ in typical models of quark matter)
then large strangelets
are unstable against fission into smaller ones
\cite{Jaikumar:2005ne,Alford:2006bx,Alford:2008ge}, and the energetically
preferred state is a crystal of strangelets: a mixed phase consisting of
nuggets of positively-charged strange matter in a neutralizing background of
electrons.
In this ``low surface tension'' scenario, strange stars are {\em not}
self-bound: they require gravitational attraction to bind the strangelets.
Stars made of strange matter are then qualitatively
similar to those made of nuclear matter:
in each case the mass-radius relation has two branches, one compact
and the other diffuse.
For nuclear matter, the compact branch contains neutron stars, which
consist of gravitationally bound nuclear matter, with an outer
crust that is a crystal of nuclei in a background of
electrons; the diffuse branch contains white
dwarfs, which are a gravitationally bound cold plasma of
nuclei (ions) and electrons, forming, at sufficiently low temperature,
a crystalline structure
(see, e.g., \cite{Potekhin:2000}).
For strange matter with a low surface tension, there are similarly two
branches. The compact
branch contains strange stars with a crust that consists of
strangelets in a background of electrons; this ``strangelet crystal crust''
was studied in Ref.~\cite{Alford:2008ge}. In this paper we
study the diffuse branch, which has no core of uniform quark matter: these
stars consist entirely of strangelets in a background of degenerate electrons,
so by analogy with white dwarfs we call them strangelet dwarfs.
The strangelet-crystal phase is a charge-separated phase.
Charge separation is favored by the internal energy
of the phases involved, because a neutral phase is always
at a maximum of the free energy with respect to the electrostatic
potential (see \cite{Ravenhall:1983uh,Glendenning:1992vb};
for a pedagogical discussion see \cite{Alford:2004hz}).
The domain structure is determined by competition between
surface tension (which favors large domains) and electric field
energy (which favors small domains). Debye screening plays
a role in determining the domain structure,
because it redistributes the electric charge, concentrating it
in the outer part of the quark matter domains and the inner part
of the surrounding electron gas, and thereby modifying the internal energy
and electrostatic energy contributions.
Our parameterization \eqn{eqn:EoS}
of the electrostatic properties of quark matter
is generic, but is not appropriate
for strangelets in the color-flavor locked (CFL) phase \cite{Alford:1998mk},
which is a degenerate
case requiring separate treatment (see Sec.~\ref{sec:params}).
To obtain the $M(R)$ relation of strangelet dwarfs,
we solve the Tolman Oppenheimer Volkoff
equation \cite{Tolman:1939jz,Oppenheimer:1939ne}, using
the equation of state of the mixed phase. We obtain the equation of state
by assuming that the strangelet lattice can be divided into unit cells
(``Wigner-Seitz cells'') and calculating the pressure of a
cell as a function of its energy density.
Our approach is similar to that used in previous studies of
the strangelet crystal \cite{Alford:2006bx,Alford:2008ge} (except that
in this paper we include electron mass effects)
and in studies of mixed phases
of quark matter and nuclear matter in the interior of neutron stars
\cite{Maruyama:2007ey}.
The main assumptions that we make are:\\
1) We assume that the strangelets in the plasma form a regular lattice
of Wigner-Seitz cells, which we treat as rotationally invariant
(spherical). In reality the cells will be unit cells of some
regular lattice. We do not consider lower-dimensional structures
(rods or slabs) because in Ref.~\cite{Alford:2008ge} we found that such
structures were never energetically favored.\\
2) Within each Wigner-Seitz cell we use a Thomas-Fermi approach,
solving the Poisson equation to obtain the charge distribution,
energy density, and pressure. This is incorrect for
very small strangelets, where the energy level structure
of the quarks becomes important \cite{Madsen:1994vp,Amore:2001uf}.
\\
3) We treat the interface between quark matter and the vacuum as a
sharp interface which is characterized
by a surface tension. We assume there is no charge localized on
the surface.
(Thus we neglect any surface charge that might arise
from the reduction of the density
of states of strange quarks at the surface
\cite{Madsen:2000kb,Madsen:2001fu,Madsen:2008bx,Oertel:2008wr}.)
\\
4) We neglect the curvature energy of a
quark matter surface \cite{Christiansen:1997rc,Christiansen:1997vt},
so we do not allow for ``Swiss-cheese'' mixed phases, in which
the outer part of the Wigner-Seitz cell is filled with quark matter, with
a cavity in the center, for which the curvature
energy is crucial.
\\
5) We work at zero temperature.
In our calculations we use units $\hbar=c=\epsilon_0=1$,
so $\alpha=e^2/(4\pi)\approx 1/137$.
\section{Phenomenological description of quark matter}
\label{sec:characterization}
We use the fact that in most phases of quark matter
the chemical potential for negative electric charge $\mue$
is much less than the chemical potential for quark number $\mu$.
This allows us to write down a model-independent parameterization of
the quark matter equation of state, expanded
in powers of $\mue/\mu$ \cite{Alford:2006bx},
\begin{equation}
\pQM(\mu,\mue) \approx
p_0(\mu)-\nQ(\mu)\mue + \half\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q(\mu) \mue^2+\ldots
\label{eqn:EoS}
\end{equation}
Note that the contribution of electrons to the pressure
of quark matter is ${\cal O}(\mue^4)$, and is neglected.
This is a very good approximation for small
strange quark mass, which corresponds to
small $\nQ$.
(For the largest value of $\nQ$
that we study, $\mue$ in neutral quark matter
is close to 100 MeV, and the assumption is still reasonable.)
As noted in Sec.~\ref{sec:intro}, we assume that
the interface between quark matter and vacuum has a
surface tension $\sigma$, and we neglect any curvature energy.
The quark density $n$ and the electric charge density
$q^{\phantom{y}}_{\rm QM}$
(in units of the positron charge) are
\begin{equation}
n = \frac{\partial \pQM}{\partial \mu},\qquad
q^{\phantom{y}}_{\rm QM} = -\frac{\partial \pQM}{\partial \mue}
= \nQ - \raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q\mue \ .
\label{charges}
\end{equation}
So in uniform neutral quark matter the electron chemical
potential is $\mue^{\rm neutral}=\nQ/\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q$. Eq.~\eqn{eqn:EoS}
is a generic parametrization if
$\mue^{\rm neutral}\ll\mu$, which is typically the case in three-flavor
quark matter.
The bag constant enters in $p_0(\mu)$, and we will fix it by
requiring that the first-order
transition between neutral quark matter and the vacuum
occur at quark chemical potential $\mu_{\rm crit}$,
i.e.~$p(\mu_{\rm crit},\mue^{\rm neutral})=0$.
Because we
are assuming that the strange matter hypothesis is valid, we require
$\mu_{\rm crit} \lesssim 310$~MeV, since at $\mu\approx 310$~MeV there is
a transition from vacuum to neutral nuclear matter.
In this article we will typically use $\mu_{\rm crit}=300~{\rm MeV}$.
The value of $\mu$ inside our quark matter lumps will always be very close
to $\mu_{\rm crit}$, so we can also expand in powers of $\mu-\mu_{\rm crit}$, and write
\begin{equation}
\begin{array}{rcl}
\pQM(\mu,\mue) &\approx&\displaystyle
n\,(\mu-\mu_{\rm crit}) + \half\chi\,(\mu-\mu_{\rm crit})^2 \\
&& + \frac{n_Q^2}{2\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q} -\nQ\mue + \half \raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q\mue^2 \ .
\end{array}
\label{generic_EoS}
\end{equation}
A quark matter equation of state can then be expressed in terms of
6 numbers: $\mu_{\rm crit}$,
the charge density $\nQ$ and charge susceptibility $\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q$
evaluated at $\mu=\mu_{\rm crit}$,
the quark number density $n$ and susceptibility $\chi$
evaluated at $\mu=\mu_{\rm crit}$, and the surface tension $\sigma$.
We will restrict ourselves
to values of the surface tension that are below the critical value
\cite{Alford:2006bx}
\begin{equation}
\begin{array}{rcl}
\si_{\rm crit} &=&\displaystyle 0.1325 \,
\frac{n_Q^2 \lambda_D}{\chi_Q}
=0.1325\,\frac{n_Q^2}{\sqrt{4\pi\alpha}\chi_Q^{3/2}},
\end{array}
\label{sicrit_result}
\end{equation}
where $\lambda_D$ is the Debye screening length in quark matter
\begin{equation}
\lambda_D = \frac{1}{\sqrt{4\pi\alpha\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q}} \ .
\label{Debye_length}
\end{equation}
If the surface tension is larger than $\si_{\rm crit}$
then the energetically favored structure at low pressure
will not be a strangelet crystal, and there will be
no strangelet dwarfs.
Rough estimates of surface tension from the
bag model are in the range $4$ to $10~{\rm MeV}/{\rm fm}^2$
\cite{Berger:1986ps,PhysRevC.44.566.2}, and
for typical models of quark matter,
$\si_{\rm crit}$ is of order $1$ to
$10~{\rm MeV}/{\rm fm}^2$ \cite{Alford:2006bx}, so it is reasonable to
explore the possibility that strange quark matter could have a surface tension
below $\si_{\rm crit}$.
\subsection{Specific equations of state}
\label{sec:specific}
When we show numerical results we will need to vary $\nQ$ and $\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q$
over a range of physically reasonable values. To give a rough idea
of what values are appropriate, we consider the example of
non-interacting three-flavor quark matter, for which $\nQ$ and $\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q$
become functions of $\mu$ and the strange quark mass
$m_s$, while $p_0$ is in addition
a function of the bag constant $B$. Expanding to lowest non-trivial
order in $m_s$,
\begin{equation}
\begin{array}{rcl}
p_0(\mu) &=& \displaystyle \frac{9\mu^4}{12\pi^2} -B \ ,\\[2ex]
\nQ(\mu,m_s) &=& \displaystyle \frac{m_s^2\mu}{2\pi^2} \ ,\\[2ex]
\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q(\mu,m_s) &=& \displaystyle \frac{2\mu^2}{\pi^2}.
\label{unpaired}
\end{array}
\end{equation}
We emphasize that these expressions are simply meant to give a rough
idea of reasonable physical values for $\nQ$ and $\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q$. Our treatment
does not depend on an expansion in powers of $m_s$.
To tune the transition between neutral quark matter and the vacuum
so it occurs at $\mu=\mu_{\rm crit}$ (see previous subsection), we set $B$ so that
$p_0(\mu_{\rm crit})=\half n_Q^2(\mu_{\rm crit})/\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q(\mu_{\rm crit})$.
In the regions between lumps of strange matter, we will assume that
there is a degenerate electron gas, whose pressure, and charge
density in units of $e$, are
\begin{equation}
\begin{array}{rcl}
p_{e^{\!-}}(\mue) &=& \displaystyle \frac{1}{24\pi^2}\biggl(
( 2 k_{Fe}^2 - 3m^2) k_{Fe} \mue \\
&&\displaystyle +\ 3 m^4 \ln\Bigl(\frac{k_{Fe}+\mue}{m}\Bigr) \biggr) \ ,\\
q_{e^{\!-}}(\mue) &=& \displaystyle -\frac{1}{3\pi^2}k_{Fe}^3 \ .
\end{array}
\label{electrons}
\end{equation}
where $\mue^2 = k_{Fe}^2 + m_e^2$.
Note that at low pressures this is
more accurate than the electron gas equation of state
used in Ref.~\cite{Alford:2008ge}, where the electron mass was set to zero.
\section{Equation of state of strangelet crystal}
\label{wigner-seitz}
\subsection{Wigner-Seitz cell}
Following the approach of \cite{Alford:2008ge}, we analyze a
spherical Wigner-Seitz cell of radius ${R_{\rm cell}}$, with a
sphere of quark matter at the center of radius $R$.
We use the Thomas-Fermi approximation to calculate $\mue(r)$,
\begin{equation}
\nabla^2 \mue(r) = -4 \pi \alpha q(r) \ ,
\label{Poisson}
\end{equation}
where $q(r)$ is the electric charge density in units of the positron charge $e$,
and $\mue$ is the electrostatic potential divided by $e$.
The boundary conditions are that there is no electric field in the
center of the cell (no $\delta$-function charge there), and
no electric field at the edge of the cell (the cell is electrically neutral),
\begin{equation}
\frac{d \mue}{d r}(0) = 0 \ , \qquad
\frac{d \mue}{d r}(R_{\rm cell}) = 0 \ .
\label{BC}
\end{equation}
We also need a matching condition at the edge of the quark matter.
Since we assume that no charge is localized on the surface, we
require continuity of $\mue$ and its first derivative (the electric field)
at $r=R$.
The value of $\mu$ inside the strange matter will be slightly
different from $\mu_{\rm crit}$ because the surface
tension compresses the droplet.
To determine the value of $\mu$, we require the pressure discontinuity
across the surface of the strangelet to be balanced by the surface tension:
\begin{equation}
\pQM(\mu,\mue(R)) - p_{e^{\!-}}(\mue(R)) = \frac{2\sigma}{R} \ .
\label{pressure_disc}
\end{equation}
Once these equations are solved, we can obtain the
equation of state of matter made of such cells.
The total energy of a cell is
\begin{equation}
\begin{array}{rcl}
E &= & \displaystyle 4\pi \int_0^R \! r^2 dr \, \Bigl(
\mu n(\mue) - \half \mue q^{\phantom y}_{\rm QM}(\mue)
- \pQM(\mu,\mue)\Bigr) \\[3ex]
&+& \displaystyle 4\pi \int_R^{{R_{\rm cell}}} \!r^2 dr \,
\Bigl( - \half \mue q_{e^{\!-}}(\mue) - p_{e^{\!-}}(\mue) \Bigr) \\[3ex]
&+& \displaystyle 4\pi R^2\, \sigma \ ,
\end{array}
\label{energy}
\end{equation}
The $-\half \mue q$ terms in \eqn{energy}
come from combining $-\mue q$ (from the relationship
between energy density and pressure) with the electric field energy density
$+\half \mue q$.
The pressure of the cell is simply the pressure of the electrons
at the edge of the cell,
\begin{equation}
p_{\rm cell} = p_{e^{\!-}}\bigl(\mue({R_{\rm cell}})\bigr) \ .
\label{pext}
\end{equation}
The total number of quarks is
\begin{equation}
N = 4\pi \int_0^R \! r^2 dr \, n(\mu,\mue) \ .
\end{equation}
The volume of the cell is $V= (4/3)\pi{R_{\rm cell}}^3$.
By varying $R$ and ${R_{\rm cell}}$ we generate a two-parameter family of
strangelets. However, there is really only a single-parameter
family of physical configurations, parameterized by the external
pressure $p_{\rm cell}$. On each line of constant $p_{\rm cell}$ in the
$(R,{R_{\rm cell}})$ parameter space, we must minimize the enthalpy per quark,
\begin{equation}
h = \frac{E + p_{\rm cell} V}{N} \ ,
\label{eq:enthalpy}
\end{equation}
to find the favored value of $R$ and ${R_{\rm cell}}$.
We assume zero temperature so $h$ is also the
Gibbs free energy per quark.
We now have a well-defined way to obtain the equation of state
of the mixed phase of quark matter, namely
the energy density $\varepsilon=E/V$ as a function
of the pressure $p_{\rm cell}$.
\subsection{Numerical solution}
\label{sec:exact}
Inside the quark matter, the solution to the Poisson equation \eqn{Poisson}
that obeys the boundary condition at the origin is
\begin{equation}
\mu_{e}(r)=\frac{n_Q}{\chi_Q}+
\frac{A}{r \lambda_D} \sinh(\frac{r}{\lambda_D}) \ ,
\label{mue-qm}
\end{equation}
where $A$ will be determined by matching conditions.
In the degenerate electron gas region outside the strange matter,
from \eqn{electrons} and \eqn{Poisson} the
Poisson equation becomes
\begin{equation}
\nabla^2 \mue(r) = \frac{4\alpha}{3\pi} (\mue^2-m_e^2)^{3/2} \ ,
\label{eqn:Poisson-vac}
\end{equation}
which must be solved numerically. For a given value of $A$ we
find from \eqn{mue-qm} the value and slope of $\mue(r)$ at $r=R$,
and use these as initial values to propagate $\mue(r)$ out to
$r={R_{\rm cell}}$ using \eqn{eqn:Poisson-vac}.
We vary $A$ until we obtain a solution that
obeys the boundary condition of no electric
field at the edge of the cell.
\subsection{Low-pressure approximations}
\label{sec:lowp}
If the pressure is not too high, the strangelet crystal consists of large
Wigner-Seitz cells (${R_{\rm cell}}\gg R$).
In this regime one can obtain approximate analytic expressions for the equation
of state of the crystal by assuming that the electrons have a roughly
constant density outside the strangelet. We give these expressions below,
and in later sections we use them to calculate mass radius relations
for large strangelet dwarf stars.
However, we expect these approximations break down at ultra-low
pressures, when the cell size becomes so
large that screening cannot be ignored, and the electrons are
clumped around the strangelets, forming atoms, rather then being
roughly uniformly distributed between the strangelets. This will
happen when ${R_{\rm cell}}$ approaches the Bohr radius
$a_0=1/(\alpha m_e)$, i.e. when $p_{\rm cell} \lesssim \alpha^5 Z^{5/3} m_e^4
\approx (10^{-12} {\rm MeV}^4) Z^{5/3}$. At these ultra-low pressures
one should use an atomic matter equation of state: we do not do this,
since we expect it will only affect a very small surface layer of the star,
without any appreciable effect on the mass-radius relationship.
The equation of state $\varepsilon(p_{\rm cell})$ is found by writing
the energy density $\varepsilon$ of the cell and its pressure as a function
of the size of the cell. For now we will
treat the size $R$ and charge $Z$ of the central strangelet as unknowns;
later we will estimate their values.
Since the pressure inside a large cell is very low the energy density
of the quark matter is approximately $n \mu_{\rm crit}$, so
\begin{equation}
\varepsilon \approx n \mu_{\rm crit} \frac{R^3}{{R_{\rm cell}}^3} \ .
\label{e-bigcell}
\end{equation}
To obtain the pressure at the edge of the cell we need to estimate the density
distribution of the electrons outside the strangelet.
\subsubsection{Constant potential approximation}
\label{sec:constapprox}
The simplest approximation is to ignore screening, taking the
electron Fermi momentum $k_{Fe}$ to be
independent of $r$ outside the strangelet
(Sec.~I of Ref.~\cite{Salpeter:1961zz}).
Imposing neutrality of the cell fixes
the Fermi momentum of the electrons,
\begin{equation}
k_{Fe}^3 = \frac{9\pi Z}{4{R_{\rm cell}}^3} \ .
\label{kFe-const}
\end{equation}
Using \eqn{e-bigcell}, we obtain the equation of state $\varepsilon(p_{\rm cell})$
of the strangelet crystal
\begin{equation}
\varepsilon \approx \mu_{\rm crit} n \frac{ 4 (k_{Fe} R)^3}{9\pi Z}
\label{eos-const}
\end{equation}
where we use \eqn{electrons} to relate the
electron Fermi momentum to $p_{\rm cell}$.
Because the constant potential approximation gives a fairly simple expression
we can use it to understand how the strangelet crystal EoS
depends on the parameters of the quark matter EoS, and hence how the
$M(R)$ curve for strangelet dwarf stars depends on those parameters.
Note that in \eqn{eos-const} the dependence of the energy density
on the pressure is via a universal and monotonically
increasing function $k_{Fe}(p)$;
dependence on the quark matter parameters
enters via the factor that multiplies this function. To make the
dependence on quark matter parameters explicit we
use results for $R$ and $Z$ from Sec.~\ref{sec:RZ} below,
and rewrite \eqn{eos-const} for the EoS of the strangelet crystal as
\begin{equation}
\begin{array}{rcl}
\varepsilon(p_{\rm cell}) &\sim& S\, \bigl(k_{Fe}(p_{\rm cell})\bigr)^3 \ , \\[1ex]
S &=&\displaystyle \frac{\mu_{\rm crit} n}{3 \pi^2 \nQ \xi(x_0(\bar\sigma))} \ ,
\end{array}
\label{eps-approx}
\end{equation}
where all dependence on the quark matter parameters comes through
the prefactor $S$, which has units of energy.
$S$ can be explicitly obtained using \eqn{sigbar-def},
\eqn{R-approx}, and \eqn{Z-strangelet} for the $\xi$ function.
One could informally think of
$S$ as a ``softness'' parameter of the
strangelet crystal EoS: as $S$ increases, the pressure becomes a more
slowly-rising function of energy density.
We expect that softer equations of state
will yield smaller stars with lower maximum masses.
In Table~\ref{tab:S-values} we give the value of $S$ for a range of
values of the parameters of the underlying quark matter EoS.
At low enough pressures, the electrons become nonrelativistic.
Then $p_{\rm cell} \approx k_{Fe}^5/(15\pi^2m_e)$, and \eqn{eos-const} simplifies to
an analytic expression for the equation of state,
\begin{equation}
\varepsilon_{NR} \approx \frac{4 R^3}{3 Z}
\Bigl(\frac{125 \pi}{9} m_e^3\Bigr)^{1/5} n \mu_{\rm crit} p_{\rm cell}^{3/5}
\label{eos-const-NR}
\end{equation}
This is a reasonable approximation when $k_{Fe}\lesssim m_e$,
i.e. when $p_{\rm cell} \lesssim m_e^4/(24\pi^2) \approx 0.0003\,{\rm MeV}^4$.
However, as we will see below, at the very lowest pressures the
constant potential approximation becomes inaccurate.
\subsubsection{Coulomb potential approximation}
\label{sec:coulapprox}
We can improve on the constant potential approximation by including the
Coulomb energy of the electrons in the calculation of the pressure.
The equation of state is still given by \eqn{eos-const}, but now
the relationship between $p_{\rm cell}$ and $k_{Fe}$ is modified
by the addition of a Coulomb energy term
(Ref.~\cite{Salpeter:1961zz},~(5)), yielding
\begin{equation}
p_{\rm cell} = p_{e^-}
- \frac{\alpha}{5} \Bigl(\frac{Z^2}{18\pi^7}\Bigr)^{1/3} k_{Fe}^4 \ .
\label{eos-Coul}
\end{equation}
Unlike the constant potential approximation, the Coulomb potential
approximation gives an energy density that goes to a non-zero value
at zero pressure,
\begin{equation}
\varepsilon_{\rm Coul}(0) = \mu_{\rm crit}\, n \frac{2 Z (\alpha m_e R)^3}{3\pi^2} \ .
\label{eps0-Coul}
\end{equation}
Comparing with \eqn{e-bigcell} we see that this corresponds to
the energy of cells with size of order $1/(\alpha m_e) \sim 10^{-10}\,{\rm m}$.
This is the energy density of a lattice of zero-pressure atomic matter
with strangelets in place of nuclei, which is a reasonable guess
for the low-pressure configuration of strangelets. We will therefore
use the Coulomb approximation as the low-pressure extension of
our equation of state.
As we will see, this leads to a ``planet'' branch
in the mass-radius relation for configurations of strange matter.
\subsubsection{Radius and charge of strangelet at low pressure}
\label{sec:RZ}
The low-pressure approximation expressions given above depend on the
size $R$ and charge $Z$ of the strangelet at the center of a
large cell. This is approximately an isolated strangelet,
whose radius can be calculated by
minimizing the isolated strangelet free energy given in eqn (25) of
Ref.~\cite{Alford:2006bx},
\begin{equation}
\overline{\Delta g}(x) = -\frac{3}{2}\frac{x - \tanh x}{x^3}
+ \frac{3\bar\sigma}{x}\ ,
\label{Delta_g_dimless}
\end{equation}
where $x$ is the radius of the strangelet in units of $\lambda_D$,
and
\begin{equation}
\bar\sigma = \frac{\sigma}{4\pi\alpha n_Q^2 \lambda_D^3} \ .
\label{sigbar-def}
\end{equation}
So the strangelet radius $R$ as a function of the parameters of the quark matter
equation of state is
\begin{equation}
R = x_0 \lambda_D,\quad \mbox{where}\quad
\frac{d \overline{\Delta g}}{dx}(x_0) = 0 \ .
\label{R-strangelet}
\end{equation}
We are interested in values of $\bar\sigma$ up to 0.13, since for higher
surface tension the strangelet crystal is
no longer stable \cite{Alford:2006bx}.
An approximate expression for the solution to \eqn{R-strangelet},
accurate to about 0.2\% for $\bar\sigma\lesssim 0.13$, is
\begin{equation}
x_0^{\rm approx} = \biggl(\frac{15\bar\sigma}{2}\biggr)^{1/3}
+ \frac{2.174 \,\bar\sigma}{1-3.982\,\bar\sigma} \ ,
\label{R-approx}
\end{equation}
where the first term is the leading-order analytic expression
for $x_0$ in the limit of small $\bar\sigma$.
The charge $Z$ of the central strangelet is given by eqn.~(17) of
Ref.~\cite{Alford:2006bx}, which can be written
\begin{equation}
\begin{array}{rcl}
Z &\approx&\displaystyle \frac{4}{3}\pi R^3 \nQ \,\xi(R/\lambda_D) \ , \\[2ex]
\xi(x) &\equiv&\displaystyle \frac{3}{x^3}(x-\tanh x) \ ,
\end{array}
\label{Z-strangelet}
\end{equation}
where $\xi$ is a correction for the effects of screening inside the
quark matter; it is an even function with $\xi(0)=1$.
\section{Numerical results}
\label{sec:results}
\subsection{Range of parameters studied}
\label{sec:params}
Our assumption that the strange matter hypothesis is valid
requires that $\mu_{\rm crit}$ must be less than
the quark chemical potential of nuclear matter, about 310~MeV, so we
fix $\mu_{\rm crit}=300~{\rm MeV}$. The value of $\mu$ inside our strange matter lumps
will always be within a few MeV of $\mu_{\rm crit}$, because if the surface tension
is small enough to favor the strangelet crystal
it will not cause significant compression.
We will perform calculations for $\lambda_D=4.82~{\rm fm}$ and $\lambda_D= 6.82~{\rm fm}$,
corresponding to $\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q \approx 0.2 \mu_{\rm crit}^2$ (appropriate for
unpaired quark matter \eqn{unpaired}) and $\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q\approx 0.1 \mu_{\rm crit}^2$
(appropriate for 2SC quark matter \cite{Alford:2006bx}).
Typical values of $\nQ$ will be around $0.05\mu_{\rm crit} m_s^2$
\eqn{unpaired}, and a reasonable range would correspond to varying
$m_s$ over its physically plausible range, from about 100 to 300 MeV.
(To have strange matter in the star, $m_s$ must be less than $\mu_{\rm crit}$.)
In this paper we use $\nQ=0.0445$, $0.0791$, and
$0.124~{\rm fm}^{-3}$,
which would correspond to $m_s=150$, 200, and 250~MeV in
\eqn{unpaired}.
There is
another widely-discussed phase of quark matter, the color-flavor locked
(CFL) phase, but it is a degenerate case where $\nQ=\raisebox{0.2ex}{$\chi$}^{\phantom{y}}_Q=0$. CFL strangelets
have a surface charge, but it does not arise from the mechanism studied
here, Debye screening, and has a different dependence on the size
of the strangelet \cite{Madsen:2001fu}. We hope to study CFL strangelet matter
in a separate work.
\begin{figure}[htb]
\includegraphics[width=\hsize]{EoS_plot}
\caption{
Equation of state of the mixed phase (strangelet crystal) for strange matter
with $\mu_{\rm crit}=300\,{\rm MeV}$,
$\lambda_D=6.82\,{\rm fm}$, $\protect\nQ=0.0791\,{\rm fm}^{-3}$,
$\sigma=1.0\,{\rm MeV}{\rm fm}^{-2}$.
The dots were obtained numerically following the procedure of
Sec.~\ref{sec:exact}. The solid line is the Coulomb-potential
approximation (Sec.~\ref{sec:coulapprox}).
The dashed line is the non-relativistic electron
(ultra-low pressure) limit \eqn{eos-const-NR}.
Above $p\approx 20000\,{\rm MeV}^4$, uniform quark matter becomes favored
over the mixed phase.
}
\label{fig:eos}
\end{figure}
\begin{figure}[htb]
\includegraphics[width=\hsize]{EoS_ratio_plot}
\caption{
Equation of state of the mixed phase for the same parameters as
in Fig.~\ref{fig:eos}, zoomed in on the low pressure region, and with
the energy density divided by $p^{0.6}$.
The dots were obtained numerically following
the procedure of Sec.~\ref{sec:exact}. The Coulomb-potential
approximation (Sec.~\ref{sec:coulapprox}) is the most accurate,
followed by the constant-potential approximation (Sec.~\ref{sec:constapprox}),
and then the non-relativistic electron
approximation \eqn{eos-const-NR}.
}
\label{fig:eos_ratio}
\end{figure}
\begin{figure}[tbh]
\includegraphics[width=\hsize]{MR_full_plot}
\caption{
The full mass-radius curve for stars made of quark matter with the equation
of state plotted in Fig.~\ref{fig:eos}, using the Coulomb approximation
\eqn{eos-Coul} to extrapolate to lower pressures.
The compact branch contains strange stars
with a strangelet crystal crust. The diffuse branch contains stars
consisting entirely of strangelet crystal matter. Solid lines
represent configurations that are stable; stability of the other branches
is discussed in the text.
}
\label{fig:MR-full}
\end{figure}
\begin{figure}
\includegraphics[width=\hsize]{MR_plot_varysigma}
\caption{
Mass-radius relation for strangelet dwarfs made of strangelet crystal
matter, comparing different approximations to the equation of state.
Upper (blue) curves are for the same parameters as in Figs.~\ref{fig:eos}
and \ref{fig:eos_ratio}. Lower (red) curves are for
a larger surface tension, $\sigma=3\,{\rm MeV}{\rm fm}^{-2}$.
The dots were obtained using the full numerical
equation of state (Sec.~\ref{sec:exact}). The solid lines
use the Coulomb-potential approximation (Sec.~\ref{sec:coulapprox}), and the
dashed lines use the constant-potential approximation \eqn{eos-const}.
}
\label{fig:MR-varysigma}
\end{figure}
\begin{figure}
\includegraphics[width=\hsize]{MR_plot_varynQ}
\caption{
Mass-radius relation for strangelet dwarfs made of strangelet crystal
matter, comparing different approximations to the equation of state.
}
\label{fig:MR-varynQ}
\end{figure}
\begin{table*}
\setlength{\tabcolsep}{1em}
\begin{tabular}{cccccccc}
$\lambda_D$ & $n_Q$ & $\si_{\rm crit}$
& \multicolumn{4}{c}{Softness prefactor $S$ (MeV) at} \\
(fm) & (fm${}^{-3}$) & (MeV fm${}^{-2}$) & $\sigma\!=\!0.3$ & $\sigma\!=\!1.0$
& $\sigma\!=\!3.0$ & $\sigma\!=\!10.0$ & \\[1ex]
\hline
4.82 & 0.0445 & 0.533 & 345 & -- & -- & -- \\
4.82 & 0.0791 & 1.69 & 158 & 202 & -- & -- \\
4.82 & 0.124 & 4.12 & 94 & 104 & 140 & -- \\[1ex]
6.82 & 0.0445 & 1.51 & 280 & 367 & -- & -- \\
6.82 & 0.0791 & 4.8 & 146 & 161 & 206 & -- \\
6.82 & 0.124 & 11.6 & 90 & 95 & 105 & 155 \\[1ex]
\hline
\end{tabular}
\caption{
Softness prefactor \eqn{eps-approx} of the strangelet crystal for various
quark matter equation of state.
The first two columns, $\lambda_D$ and $\protect\nQ$,
specify the quark matter equation of state \eqn{generic_EoS}
(via \eqn{Debye_length}). The third column gives the maximum surface tension
for which a strangelet crystal will occur \eqn{sicrit_result}.
The last four columns give the softness prefactor $S$ for different values
of the surface tension $\sigma$ (given in MeV\,fm${}^{-2}$)
of the interface between quark matter and vacuum.
}
\label{tab:S-values}
\end{table*}
\subsection{Testing approximations to the equation of state}
In Fig.~\ref{fig:eos} we show the equation of state for the
strangelet crystal, for critical quark chemical potential $\mu_{\rm crit}=300\,{\rm MeV}$,
quark matter screening distance $\lambda_D=6.82\,{\rm fm}$,
quark charge density parameter $\nQ=0.0791\,{\rm fm}^{-3}$,
and quark matter surface tension $\sigma=1.0\,{\rm MeV}{\rm fm}^{-2}$.
The dots were obtained numerically following the procedure of
Sec.~\ref{sec:exact}. The solid line is the Coulomb-potential
approximation (Sec.~\ref{sec:coulapprox}). On this
plot the constant potential
approximation (Sec.~\ref{sec:constapprox}) line would be indistinguishable
from the Coulomb-potential line, so we do not show it.
The dot-dashed line is the non-relativistic electron
(ultra-low pressure) limit \eqn{eos-const-NR} of the constant potential
approximation.
Above $p\approx 20000\,{\rm MeV}^4$, uniform quark matter becomes favored
over the mixed phase.
On this very expanded logarithmic scale, the Coulomb approximation
appears reasonably
accurate up to pressures of order 1\,MeV.
To achieve more discrimination between the different approximations, we
show in Fig.~\ref{fig:eos_ratio} a magnified version of the low-pressure
end of the plot in Fig.~\ref{fig:eos}, where we have divided out the
non-relativistic scaling of the energy density, $\varepsilon \sim p^{3/5}$.
We can see that, down to the lowest pressures for which we can
perform the numerical Wigner-Seitz calculation of the equation of state,
the Coulomb approximation gives the most accurate semi-analytic
approximation, although the constant potential approximation is
accurate to within about 10\%.
We then have to decide which approximation to use for lower
pressures, where numerical calculations are not available.
In the low-pressure limit, the Coulomb approximation to $\varepsilon(p)$ tends to a
fixed value, while the constant and nonrelativistic approximations to $\varepsilon(p)$
tend to zero as $p^{3/5}$. So in Fig.~\ref{fig:eos_ratio} the Coulomb
approximation will diverge at $p\ll 10^{-5}\,{\rm MeV}^4$, while the constant and
nonrelativistic approximations will tend to the same constant value.
As discussed in Sec.~\ref{sec:coulapprox}, it seems reasonable to expect that at
the lowest pressures there will be a crystal of ``strange atoms'', each
consisting of electrons bound to a strangelet, and the Coulomb approximation
gives a reasonable estimate of the energy density of such matter,
so at low pressure we will
use the Coulomb approximation.
\subsection{Mass-radius relation of strange stars}
In Fig.~\ref{fig:MR-full} we show
the full mass-radius curve for stars made of quark matter with the equation
of state plotted in Fig.~\ref{fig:eos}.
The compact branch contains strange stars
with a strangelet crystal crust. The diffuse branch contains stars
consisting entirely of strangelet crystal matter. It includes
two segments: the lighter one is planets of dilute
strange matter whose the mass increases with radius.
This joins to the strangelet dwarf
branch where the mass decreases with radius as the strangelet crystal
is compressed
by the pressure due to gravity. We use the numerically
calculated equation of state (Sec.~\protect\ref{wigner-seitz}) except that
at very low pressure (the planetary branch) the Wigner-Seitz cells
become so large that our numerical methods break down, so as
discussed in Sec.~\ref{sec:coulapprox} we
use the Coulomb approximation \eqn{eos-Coul} to extrapolate
down to zero pressure.
Fig.~\ref{fig:MR-full} shows the whole $M(R)$ curve, not all of which
corresponds to stable configurations.
The usual stability criterion for stars \cite{Bardeen:1966} is that
one radial mode becomes either stable or unstable at each extremum in the
$M(R)$ function. A stable mode
becomes unstable at each extremum where
the curve bends counterclockwise as the central density increases;
a stable mode becomes unstable at each extremum where the curve bends
clockwise as the central density increases. However,
Glendenning~et.al.~\cite{Glendenning:1994sp}
report that at some extrema there is no change in stability: the squared
frequency of one of the fundamental radial modes may touch zero, but
not change sign.
We defer a detailed study of the stability of radial modes of
strange stars to future work, and in Fig.~\ref{fig:MR-full} we
show as ``stable'' (solid curves) the parts of the $M(R)$ curve that both
Ref.~\cite{Bardeen:1966} and Ref.~\cite{Glendenning:1994sp} agree are stable.
We note that Ref.~\cite{Glendenning:1994sp} is a study of
stars that have a core of uniform strange matter surrounded
by a crust of nuclear matter: these are
similar to the configurations along the dashed part
of the mass-radius curve in Fig.~\ref{fig:MR-full}, where we have a core
of uniform strange matter surrounded by a crust of strangelets, with a
density discontinuity at the boundary. If Ref.~\cite{Glendenning:1994sp}'s
stability argument is correct and applicable to our stars,
then some of these configurations may also be stable.
In the remainder of this paper we will focus on the strangelet dwarf
branch, which consists of a simple crystal of stranglets with no
uniform core, so there is no controversy about the appropriate
stability criterion.
\subsection{Mass-radius relation of strangelet dwarfs}
To investigate the sensitivity of the masses and radii of strangelet dwarfs to
the parameters of the quark matter equation of state, we show in
Fig.~\ref{fig:MR-varysigma} and ~\ref{fig:MR-varynQ}
the strangelet dwarf part of the
mass-radius curve, excluding the compact and planetary branches, for
various values of the quark matter parameters.
In Fig.~\ref{fig:MR-varysigma} we explore the effects of varying
the surface tension, and we compare the different approximations to
the equation of state.
The upper curves are for the same equation of state
as was shown in Fig.~\ref{fig:eos}
and \ref{fig:eos_ratio}; the lower curves use a larger surface tension,
$\sigma=3\,{\rm MeV}{\rm fm}^{-2}$. In both cases the solid curves are obtained from the
Coulomb-potential approximation to the equation of state,
and the dashed lines are obtained from the constant-potential
approximation. The dots use the equation of state that
is obtained numerically following the procedure of
Sec.~\ref{sec:exact}, except that at very low pressures, where the
numerical calculation becomes too difficult, the Coulomb approximation
is used.
We see that, as one might have expected from Fig.~\ref{fig:eos},
using the Coulomb approximation over the entire pressure range of the
mixed phase yields reasonably accurate results.
However, as noted in Sec.~\ref{sec:constapprox},
the constant potential approximation is still useful
for gaining an understanding of how the $M(R)$ curve for strangelet dwarfs
depends on the parameters of the EoS, because in the range of
pressures that is important for strangelet dwarfs it gives
a good indication of the $M(R)$ curve.
(At ultra-low pressures,
relevant for the strange planet branch, this is no longer the case: one has
to use the Coulomb approximation instead.)
As discussed in Sec.~\ref{sec:constapprox}, the constant potential
approximation to the EoS can be written in terms of a ``softness prefactor''
$S$ \eqn{eps-approx}. To understand how the $M(R)$ curve in
Fig.~\ref{fig:MR-varysigma} changes with $\sigma$, note that
$x_0(\bar\sigma)$ is a monotonically increasing
function and $\xi(x_0)$ is a monotonically decreasing function,
so as the surface tension $\sigma$ increases at fixed values
of the other parameters, the softness prefactor $S$ of the
strangelet crystal EoS increases
(one can see this in Table~\ref{tab:S-values}). Since the
EoS is becoming softer, the $M(R)$ curve moves down
and to the left, giving smaller stars with a lower maximum mass.
In Fig.~\ref{fig:MR-varynQ} we explore the effects of varying
the charge density parameter $\nQ$ in \eqn{generic_EoS}
while keeping the other parameters constant.
As in Fig.~\ref{fig:MR-varysigma}, solid lines are for the Coulomb approximation
to the equation of state, dots are for the numerically calculated
equation of state using the Coulomb approximation to extrapolate to the
lowest pressures. We see that increasing $\nQ$ yields
heavier, larger strangelet dwarf stars. Again, this can be understood
in terms of the constant potential approximation and its softness
prefactor $S$ \eqn{eps-approx}. As $\nQ$ increases, it causes
$S$ to decrease through two effects. Firstly via
the explicit factor of $\nQ$ in the denominator of \eqn{eps-approx},
and secondly via the relationship \eqn{sigbar-def} between
$\sigma$ and $\bar\sigma$. The sensitivity of $S$ to changes in $\nQ$ can
be seen in Table~\ref{tab:S-values}: for the two values of $\nQ$
used in Fig.~\ref{fig:MR-varynQ} the values of $S$ are near the
extremes of its range in the parameter set we studied: $S\approx 345$
and $S\approx 94$ for $\nQ=0.0445$ and $\nQ=0.124$ respectively.
Consequently, the $M(R)$ curve for $\nQ=0.0445$ is characteristic of
a soft equation of state, with low radius at a given mass and a low
maximum mass, whereas the $M(R)$ curve for $\nQ=0.124$ is characteristic of
a hard equation of state, with large radius at a given mass and a high
maximum mass.
\section{Discussion}
\label{sec:discussion}
We have shown that, if the strange matter hypothesis is correct
and the surface tension of the interface between strange matter
and the vacuum is less
than a critical value \eqn{sicrit_result}, there is at least one
additional stable branch in the mass-radius relation for strange stars,
corresponding to large diffuse objects that we call ``strangelet dwarfs'',
consisting of a crystal of strangelets in a sea of electrons.
This is easily understood, since if $\sigma<\si_{\rm crit}$ then uniform strange matter
is unstable at zero pressure, and undergoes
charge separation to a crystal of positively-charged
strangelets surrounded by electrons,
just as normal matter at zero pressure is a mixed phase
consisting of droplets of
nuclear matter surrounded by electrons. Strangelet dwarfs are then
the strange matter equivalent of white dwarfs.
We emphasize that in this low-surface-tension scenario, strange matter
is {\em not} self bound. Like nuclear matter, it is only bound by
gravitational forces. Every strange star will have a strangelet crystal
crust, and strangelet dwarfs are those strange stars that are ``all crust''.
The natural production mechanism by which strangelet dwarfs might
be produced is a collision between a strange star and another compact
object. In such collisions, up to $0.03\,M_\odot$
may be ejected \cite{Bauswein:2008gx}, which is in the mass range we are
predicting for strangelet dwarfs. There are two ways a collision could
produce strangelet dwarfs. Firstly, part of the crust of the strange star
might be ejected to become a isolated object, which would be a
strangelet dwarf. Secondly, if a sufficiently light piece of
the uniform quark matter core were ejected in the collision,
it would be unable to
exist on the compact branch, and would evaporate into
a configuration on the diffuse branch. For example,
for the equation of state studied in Fig.~\ref{fig:MR-full}, the
lightest compact configuration of strange matter is $0.0055\,M_\odot$.
A lighter piece of strange matter could only exist on the diffuse branch,
and would spontaneously evaporate to become a strangelet dwarf.
Strangelet dwarfs produced by these mechanisms could then bind
gravitationally, to form heavier strangelet dwarfs.
It should be noted that our proposed mechanism for the production
of strangelet dwarfs is also a mechanism for creating a diffuse cosmic
flux of strangelets (``strangelet pollution''), which might be expected to
convert all neutron stars
to strange stars \cite{Friedman:1990qz}. Although observations of
glitches and magnetar oscillations \cite{Watts:2006hk} seem consistent
with some compact stars having nuclear matter crusts, there remains some
uncertainty. Crystalline phases of quark matter could allow strange stars
to glitch \cite{Mannarelli:2007bs}, and in our low-surface-tension scenario
strange stars have crusts that could be hundreds of meters thick
\cite{Alford:2008ge}. A cosmic flux of strangelets may seem unlikely but until
it is ruled out experimentally (as may happen soon from the AMS experiment
\cite{Sandweiss:2004bu}) it remains useful to analyze the full observational
consequences of the strange matter hypothesis.
Our analysis assumes that at any
given pressure the strangelet crystal consists of the most energetically
favorable strangelet configuration (in terms of strangelet size and charge
and cell size). However, other configurations will in general be metastable
with long lifetimes. If one compresses a piece of strangelet crystal then
the charge of the strangelets can readily change via absorption or
emission of electrons, but it is very difficult for the
quark matter to rearrange itself in to strangelets of the
now-energetically-favored size: it is more likely that
the strangelets will stay the same size and the radial density profile
of the electrons will change. The sizes of the strangelets will
be determined more by the history of the object than by the pressure.
Taking this point further, it is quite possible to have a
crystal consisting of a mixture of strangelets and ordinary nuclei,
held apart by their electrostatic repulsion but also bound together
in to a crystal by the degenerate electron gas that neutralizes them,
forming a hybrid strangelet/white dwarf star.
Detection of strangelet dwarfs requires an observation method that can find
non-luminous objects with typical masses of $10^{-5}$ to $10^{-1}\,M_\odot$
and radii in the range $500$ to $5000$\,km. An example is
gravitational microlensing surveys, such as those conducted by
the Microlensing Observations in Astrophysics (MOA) and the Optical
Gravitational Lensing Experiments (OGLE) groups, which look for
lensing events in the galactic bulge, and are capable of detecting
Jupiter-mass objects. It is intriguing that such surveys now
report the existence of an abundant population of unbound distant planetary
masses, suggesting that such objects may be twice as common as main
sequence stars \cite{Sumi:2011}. Although models of planet formation
indicate that mechanisms exist for unbinding planets through disk instabilities
and planet interactions \cite{Veras:2009}, we suggest that a possible
alternative is formation of strange dwarfs from matter ejected in strange
star mergers. One would expect that sometimes a strangelet dwarf produced
in a merger might be unable to escape the gravitational field of
the remaining compact object, and this would explain the presence of
dense planet-mass objects in the vicinity of compact stars. An example is
the millisecond pulsar PSR J1719-1438, which has a Jupiter-mass
companion whose inferred central density
($\rho > 23\,{\rm g}\,{\rm cm}^{-3}$) is
far in excess of what is expected in a planet \cite{Bailes:2011}.
We expect that in the near future further light will be cast on this question,
as microlensing surveys help us better
understand the distribution of planetary mass compact objects and
as strategies are devised to provide information about both mass and radius.
\section*{Acknowledgments}
We thank E. Agol and F. Weber for useful discussions.
This research was
supported in part by the
Offices of Nuclear Physics and High Energy Physics of the
Office of Science of the
U.S.~Department of Energy under contracts
\#DE-FG02-91ER40628,
\#DE-FG02-05ER41375,
and by the DoE Topical Collaboration
``Neutrinos and Nucleosynthesis in Hot and Dense Matter'',
contract \#DE-SC0004955.
\renewcommand{\href}[2]{#2}
\newcommand{Astrophys. J. Lett.\ }{Astrophys. J. Lett.\ }
\newcommand{Mon. Not. R. Astron. Soc.\ }{Mon. Not. R. Astron. Soc.\ }
\newcommand{Astron. Astrophys.\ }{Astron. Astrophys.\ }
\bibliographystyle{JHEP_MGA}
|
\section{Introduction}
\label{sec:intro}
Many algorithms in machine learning and other scientific computing fields rely on optimizing a function with respect to a parameter space. In many cases, the objective function being optimized takes the form of a sum over a large number of terms that can be treated as identically distributed: for instance, labeled training samples. Commonly, the problem that we are trying to solve consists of minimizing the negated log-likelihood:
\begin{equation}
f(\btheta) = -\log(p(\Y|\X;\btheta)) = -\sum_{i=1}^N\log(p(\y_i|\x_n;\btheta)) \label{eqn:objf}
\end{equation}
where $(\X,\Y)$ are our observations and labels respectively, and $p$ is the posterior probability of our labels which is modeled by a deep neural network with parameters $\btheta$. In this case it is possible to use subsets of the training data to obtain noisy estimates of quantities such as gradients; the canonical example of this is Stochastic Gradient Descent (SGD).
The simplest reference point to start from when explaining our method is Newton's method with line search, where on iteration $m$ we do an update of the form:
\begin{equation}
\theta_{m+1} = \theta_m - \alpha \H_m^{-1} \g_m, \label{eqn:newton}
\end{equation}
where $\H_m$ and $\g_m$ are, respectively, the Hessian and the gradient on iteration $m$ of the objective function~\eqref{eqn:objf}; here, $\alpha$ would be chosen to minimize~\eqref{eqn:objf} at $\theta_{m+1}$. For high dimensional problems it is not practical to invert the Hessian; however, we can efficiently approximate~\eqref{eqn:newton} using only multiplication by $\H_m$, by using the Conjugate Gradients (CG) method with a truncated number of iterations. In addition, it is possible to multiply by $\H_m$ without explicitly forming it, using what is known as the ``Pearlmutter trick''~\cite{pearlmutter1994fast} (although it was known to the optimization community prior to that; see~\cite[Chapter 8]{Nocedal2006NO}) for multiplying an arbitrary vector by the Hessian; this is described for neural networks but is applicable to quite general types of functions. This type of optimization method is known as ``truncated Newton'' or ``Hessian-free inexact Newton''~\cite{Morales00enrichedmethods}. In~\cite{byrd2011use}, this method is applied but using only a subset of data to approximate the Hessian $\H_m$. A more sophisticated version of the same idea was described in the earlier paper~\cite{martens2010deep}, in which preconditioning is applied, the Hessian is damped with the unit matrix in a Levenberg-Marquardt fashion, and the method is extended to non-convex problems by substituting the Gauss-Newton matrix for the Hessian. We will discuss the Gauss-Newton matrix and its relationship with the Hessian in Section~\ref{sec:gn}.
Our method is quite similar to the one described in~\cite{martens2010deep}, which we will refer to as Hessian Free (HF). We also multiply by the Hessian (or Gauss-Newton matrix) using the Pearlmutter trick on a subset of data, but on each iteration, instead of approximately computing $(\H_m + \lambda \I)^{-1} \g_m$ using truncated CG, we compute a basis for the Krylov subspace spanned by $\g_m, \H_m \g_m, \ldots \H_m^{K-1} \g_m$ for some $K$ fixed in advance (e.g. $K=20$), and numerically optimize the parameter change within this subspace, using BFGS to minimize the original nonlinear objective function measured on a subset of the training data. It is easy to show that, for any $\lambda$, the approximate solution to $\H_m + \lambda \I$ found by $K$ iterations of CG will lie in this subspace, so we are in effect automatically choosing the optimal $\lambda$ in the Levenburg-Marquardt smoothing method of HF (although our algorithm is free to choose a solution more general than this). We note that both our method and HF use preconditioning, which we have glossed over in the discussion above. Compared with HF, the advantages of our method are:
\begin{itemize}
\item Greater simplicity and robustness: there is no need for heuristics to initialize and update the smoothing value $\lambda$.
\item Generality: unlike HF, our method can be applied even if $\H$ (or whatever approximation or substitute we use) is not positive semidefinite.
\item Empirical advantages: our method generally seems to work better than HF in both optimization speed and classification performance.
\end{itemize}
The chief disadvantages versus HF are:
\begin{itemize}
\item Memory requirement: we require storage of $K$ times the parameter dimension to store the subspace.
\item Convergence properties: the use of a subset of data to optimize over the subspace will prevent convergence to an optimum.
\end{itemize}
Regarding the convergence properties: we view this as more of a theoretical than a practical problem, since for typical setups in training deep networks the residual parameter noise due to the use of data subsets would be far less than that due to overtraining.
Our motivation for the work presented here is twofold: firstly, we are interested in large-scale non-convex optimization problems where the parameter dimension and the number of training samples is large and the Hessian has large condition number. We had previously investigated quite different approaches based on preconditioned SGD to solve an instance of this type of optimization problem (our method could be viewed as an extension to~\cite{le2007topmoumoute}), but after reading~\cite{martens2010deep} our interest switched to methods of the HF type. Secondly, we have an interest in deep neural nets, particularly to solve problems in speech recognition, and we were intrigued by the suggestion in~\cite{martens2010deep} that the use of optimization methods of this type might remove the necessity for pretraining, which would result in a welcome simplification. Other recent work on the usefulness of second order methods for deep neural networks includes~\cite{GlorotAISTATS2010,NgICML11}.
\section{The Hessian matrix and the Gauss-Newton matrix}
\label{sec:gn}
The Hessian matrix $\H$ (that is, the matrix of second derivatives w.r.t. the parameters) can be used in HF optimization whenever it is guaranteed positive semidefinite, i.e. when minimizing functions that are convex in the parameters. For non-convex problems, it is possible to substitute a positive definite approximation to the Hessian. One option is the Fisher information matrix,
\begin{equation}
\F = \sum_i \g_i \g_i^T,
\end{equation}
where indices $i$ correspond to samples and the $\g_i$ quantities are the gradients
for each sample. This is a suitable stand-in for the Hessian because
it is in a certain sense dimensionally the same, i.e. it changes the same way under
transformations of the parameter space. If the model can be interpreted as producing
a probability or likelihood, it is possible under certain assumptions (including model
correctness) to show that close to convergence, the Fisher and Hessian matrices have
the same expected value. The use of the Fisher matrix in this way is known as
Natural Gradient Descent~\cite{Amari:1998:NGW:287476.287477}; in~\cite{le2007topmoumoute},
a low-rank approximation of the Fisher matrix was used instead.
Another alternative that has less theoretical justification
but which seems to work better in practice in the case of neural networks is the
Gauss-Newton matrix, or rather a
slight generalization of the Gauss-Newton matrix that we will now describe.
\subsection{The Gauss-Newton matrix}
The Gauss-Newton matrix is defined when we have a function (typically nonlinear) from a
vector to a vector, $f: \Re^n \rightarrow \Re^m$. Let the Jacobian of this function be
$\J \in \Re^{m \times n}$, then the Gauss-Newton matrix is $\G = \J^T \J$,
with $\G \in \Re^{n\times n}$. If the problem is least-squares on the output
of $f$, then $\G$ can be thought of as one term
in the Hessian on the input to $f$. In its application to neural-network training,
for each training example we consider the network as a nonlinear function from the
neural-network parameters $\btheta$ to the output of the network, with the neural-network input
treated as a constant. As in~\cite{schraudolph}, we
generalize this from least squares to general convex error functions by using the
expression $\J^T \H \J$, where $\H$ is the (positive semidefinite) second derivative
of the error function w.r.t. the neural network output. This
can be thought of as the part of the Hessian that remains after ignoring the
nonlinearity of the neural-network in the parameters. In the rest of this document,
following~\cite{martens2010deep} we will refer to this matrix $\J^T \H \J$ simply as
the Gauss-Newton matrix, or $\G$, and depending on the context, we may actually be
referring to the summation of this expression over a number of neural-network training
samples.
\subsection{Efficiently multiplying by the Gauss-Newton matrix}
As described in~\cite{schraudolph}, it is possible to efficiently multiply a
vector by $\G$ using a version of the ``Pearlmutter trick''; the algorithm is similar
in spirit to backprop and we give it as Algorithm~\ref{alg:gn}. Our notation and our derivation for this
algorithm differ from~\cite{pearlmutter1994fast,schraudolph}, and we will explain this briefly;
we find our approach easier to follow. The idea is this: we first imagine
that we are given a parameter set $\btheta$, and two vectors $\btheta_1$
and $\btheta_2$ which we interpret as directions in parameter space; we then
write down an algorithm to compute the scalar $s = \btheta_2^T \G \btheta_1$.
Assume the neural-network input is given and fixed;
let $\v$ be the network output, and write it as $\v(\btheta)$ to emphasize the
dependence on the parameters, and then let $\v_1$ be defined as
\begin{equation}
\v_1 = \lim_{\alpha \rightarrow 0} \frac{1}{\alpha} \v(\btheta + \alpha \btheta_1) - \v(\btheta), \label{eqn:v1}
\end{equation}
so that $\v_1 = \J \btheta_1$. We define $\v_2$ similarly. These can both
be computed in a modified forward pass through the network. Then, if $\H$ is the
Hessian of the error function in the output of the network (taken at parameter value $\btheta$),
$s$ is given by
\begin{equation}
s = \v_2^T \H \v_1, \label{eqn:s:1}
\end{equation}
since $\v_2^T \H \v_1 = \btheta_2^T \J^T \H \J \btheta_1 = \btheta_2^T \G \btheta_1$.
The Hessian $\H$ of the error
function would typically not be constructed as a matrix, but we would compute~\eqref{eqn:s:1}
given some analytic expression for $\H$. Suppose we have written down the algorithm
for computing $s$ (we have not done so here because of space constraints). Then
we treat $\btheta_1$ as a fixed quantity, but compute the derivative of $s$ w.r.t.
$\btheta_2$ (taking $\btheta_2$ around zero for convenience). This derivative equals
the desired product $\G \btheta_1$. Taking the derivative of a scalar
w.r.t. the input to an algorithm can be done in a mechanical fashion via ``reverse-mode''
automatic differentiation through the algorithm, of which neural-net backprop is a special
case. This is how we obtained Algorithm~\ref{alg:gn}.
In the algorithm we denote the derivative of $s$ w.r.t. a quantity $x$ by $\hat{x}$, i.e.
by adding a hat. We note that in this algorithm, we have a ``backward pass'' for
quantities with subscript 2,
which did not appear in the forward pass, because
they were zero (since we take $\btheta_2 = 0$) and we optimized them out.
Something to note here is that when the linearity of the last layer is softmax and the error
is negated cross-entropy (equivalently negated log-likelihood, if the label is known),
we actually view the softmax nonlinearity as part of the error function. This is a
closer approximation to the Hessian, and it remains positive semidefinite.
To explain the notation of Algorithm~\ref{alg:gn}: $\h^{(i)}$ is the input
to the nonlinearity of the $i$'th layer and $\v^{(i)}$ is the output;
$\odot$ means elementwise multiplication; $\phi^{(i)}$ is the nonlinear function of
the $i$'th layer, and when we apply it
to vectors it acts elementwise; $\W^{(1)}$ is the neural-network weights for the
first layer (so $\h^{(1)} = \W^{(1)} \v^{(0)}$, and so on); we use the subscript
$1$ for quantities that represent how quantities change when we move the parameters
in direction $\btheta_1$ (as in~\eqref{eqn:v1}).
The error function is written as ${\cal E}(\v^{(L)}, y)$ (where $L$ is the last layer),
and $y$, which may be a discrete value, a scalar or a vector, represents the
supervision information the network is trained with. Typically ${\cal E}$ would
represent a squared loss or negated cross-entropy.
In the squared-loss case, the quantity $\frac{\partial^2}{\partial \v^2} {\cal E}(\v^{(L)}, y)$
in Line~\ref{line:e} of Algorithm~\ref{alg:gn} is just the unit matrix. The other
case we deal with here is negated cross entropy. As mentioned above, we include
the soft-max nonlinearity in the error function,
treating the elements of the output layer $\v^{(L)}$ as unnormalized log probabilities. If
the elements of $\v^{(L)}$ are written as $v_j$ and we let $\p$ be the vector of
probabilities, with $p_j = \exp(v_j) / \sum_i \exp(v_i)$, then the
matrix of second derivatives is given by
\begin{equation}
\frac{\partial^2}{\partial \v^2} {\cal E}(\v^{(L)}, y) = \diag(\p) - \p \p^T .
\end{equation}
\begin{algorithm}
\caption{Compute product $\hat{\btheta}_2 = \G \btheta_1$: MultiplyG$(\btheta, \btheta_1, \x, y)$ }
\label{alg:gn}
\begin{algorithmic}[1]
\STATE \algorithmiccomment{ Note, $\btheta = ( \W^{(1)}, \W^{(2)}, \ldots )$ and $\btheta_1 = ( \W_1^{(1)}, \W_2^{(2)}, \ldots )$. }
\STATE $\v^{(0)} \gets \x$
\STATE $\v_1^{(0)} \gets {\mathbf 0}$
\FOR { $l = 1 \ldots L$ }
\STATE $\h^{(l)} \gets \W^{(l)} \v^{(l-1)}$
\STATE $\h_1^{(l)} \gets \W^{(l)} \v_1^{(l-1)} + \W_1^{(l)} \v^{(l-1)}$
\STATE $\v^{(l)} \gets \phi^{(l)}(\h^{(l)})$
\STATE $\v_1^{(l)} \gets {\phi'}^{(l)}(\h^{(l)}) \odot \h_1^{(l)}$
\ENDFOR
\STATE $\hat{\v}_2^{(L)} \gets \frac{\partial^2}{\partial \v^2} {\cal E}(\v^{(L)}, y) \v_1^{(L)}$ \label{line:e}
\FOR { $l = L \ldots 1$ }
\STATE $\hat{\h}_2^{(l)} \gets \hat{\v}_2^{(l)} \odot {\phi'}^{(l)}(\h^{(l)})$
\STATE $\hat{\v}_2^{(l-1)} \gets \left.\W^{(l)}\right.^T \hat{\h}_2^{(l)}$
\STATE $\hat{\W}_2^{(l)} \gets \hat{\h}_2^{(l)} \left.\v^{(l-1)}\right.^T$
\ENDFOR
\RETURN $\hat{\btheta}_2 \equiv \left( \hat{\W}_2^{(1)}, \ldots, \hat{\W}_2^{(L)} \right)$
\end{algorithmic}
\end{algorithm}
\section{Krylov Subspace Descent: overview}
\label{sec:overview}
Now we describe our method, and how it relates to Hessian Free (HF) optimization.
The discussion in the previous section (on the Hessian versus Gauss-Newton matrix) is orthogonal
to the distinction between KSD and HF, because either method can use any Hessian substitute, with the
proviso that our method can use the Hessian even when it is not positive definite.
In the rest of this section we will use $\H$ to refer to either the Hessian or a substitute such
as $\G$ or $\F$. In \cite{martens2010deep} and the work we describe here, these matrices are approximated using a subset of data samples.
In both HF and KSD, the whole computation is preconditioned using the diagonal of $\F$ (since this is
easy to compute); however, in the discussion below we will gloss over this preconditioning.
In HF, on each iteration the CG algorithm is used to approximately compute
\begin{equation}
\d = - (\H + \lambda \I)^{-1} \g,
\end{equation}
where $\d$ is the step direction, and $\g$ is the gradient. The step size is determined by a
backtracking line search. The value of $\lambda$ is kept updated by Levenburg-Marquardt style
heuristics. Other heuristics are used to control the stopping of the CG iterations. In addition,
the CG iterations for optimizing $\d$ are not initialized from zero (which would be the natural
choice) but from the previous value of $\d$; this loses some convergence guarantees but seems to
improve performance, perhaps by adding a kind of momentum to the updates.
In our method (again glossing over preconditioning), we compute a basis for the subspace
spanned by $\{ \g, \H \g, \ldots, \H^{K-1} \g, \d_{\mathrm{prev}} \}$, which is the Krylov
subspace of dimension $K$, augmented with the previous search direction. We then optimize the objective function over this subspace using BFGS, approximating
the objective function using a subset of samples.
\section{Krylov Subspace Descent in detail}
In this section we describe the details of the KSD algorithm, including the
preconditioning.
For notation purposes: on iteration $n$ of the overall optimization we will write
the training data set used to obtain the gradient as ${\cal A}_n$ (which is
always the entire dataset in our experiments); the set used to compute the Hessian
or Hessian substitute as ${\cal B}_n$; and the set used for BFGS
optimization over the subspace, as ${\cal C}_n$.
For clarity when dealing with multiple subset sizes, we will typically normalize all
quantities by the number of samples: that is, objective function values, gradients,
Hessians and the like will always be divided by the number of samples in the
set over which they were computed.
On each iteration we will compute a diagonal preconditioning matrix $\D$ (we omit
the subscript $n$). $\D$ is expected to be a rough approximation to the Hessian.
In our experiments, following~\cite{martens2010deep}, we set $\D$ to the diagonal
of the Fisher matrix computed over ${\cal A}_n$.
To precondition, we define a new variable $\tilde{\btheta} = \D^{1/2} \btheta$, compute the Krylov
subspace in terms of this variable, and convert back to the ``canonical'' co-ordinates.
The result is the subspace spanned by the vectors
\begin{equation}
\left\{ (\D^{-1} \H)^k \D^{-1} \g, 0\leq k < K \right\} \label{eqn:subs}
\end{equation}
We adjoin the previous step direction $\d_\mathrm{prev}$ to this, and it becomes the
subspace we optimize over with BFGS. The algorithm to compute an orthogonal
basis for the subspace, and the Hessian (or Hessian substitute) within it, is given as
Algorithm~\ref{alg:proj}.
\begin{algorithm}[h]
\caption{Construct basis $\V = \left[\v_1,\ldots,\v_{K+1}\right]$ for the subspace, and the Hessian (or substitute) $\bar{\H}$ in the co-ordinates of the subspace.}
\label{alg:proj}
\begin{algorithmic}[1]
\STATE $\v_1 \gets \D^{-1} \g$
\STATE $\v_1 \gets \frac{1}{\sqrt{\v_1^T \v_1}} \v_1$
\FOR { $k = 1 \ldots K+1$ }
\STATE $\w \gets \H \v_k$ \algorithmiccomment{If Gauss-Newton matrix, computed with Algorithm~\ref{alg:gn}.}
\IF { $k < K$ }
\STATE $\u \gets \D^{-1} \w$ \algorithmiccomment{$\u$ will be $\v_{m+1}$}
\ELSIF { $k = K$ }
\STATE $\u \gets \d_\mathrm{prev}$ \algorithmiccomment{Previous search direction; use arbitrary nonzero vector if 1st iter}
\ENDIF
\FOR { $j = 1 \ldots k$ }
\STATE $\bar{h}_{k,j} \gets \w^T \v_j$ \algorithmiccomment{Compute element of reduced-dimension Hessian}
\STATE $\u \gets \u - (\u^T \v_j) \v_j$ \algorithmiccomment{Orthogonalize $\u$}
\ENDFOR
\IF { $k \leq K$}
\STATE $\v_{k+1} \gets \frac{1}{\sqrt{\u^T \u}} \u$ \algorithmiccomment{Normalize length and set next direction.}
\ENDIF
\ENDFOR
\STATE \algorithmiccomment{Now set upper triangle of $\bar{\H}$ to lower triangle.}
\end{algorithmic}
\end{algorithm}
On each iteration of optimization, after computing the basis $\V$ with Algorithm~\ref{alg:proj}
we do a further preconditioning step within the subspace, which gives us a new, non-orthogonal
basis $\hat{\V}$ for the subspace. This step is done to help the BFGS converge faster.
\begin{algorithm}[h]
\caption{Krylov Subspace Descent}
\label{alg:ksd}
\begin{algorithmic}[1]
\STATE $\d_\mathrm{prev} \gets \e_1$ \algorithmiccomment{or any arbitrary nonzero vector}
\FOR { $n = 1, 2 \ldots$ }
\STATE \algorithmiccomment{Sample three sets from training data, ${\cal A}_n$, ${\cal B}_n$ and ${\cal C}_n$.}
\STATE $\g \gets \frac{1}{ |{\cal A}_n| } \sum_{i \in {\cal A}_n} \g_i(\btheta)$ \algorithmiccomment{ Get average function gradient over this batch. }
\STATE Set $\D$ to diagonal of Fisher matrix on ${\cal A}_n$, floored to $\epsilon$ times its maximum.
\STATE Run Algorithm~\ref{alg:proj} to find $\V$ and $\bar{\H}$ on subset ${\cal B}_n$
\STATE Let $\hat{\H}$ be the result of flooring the eigenvalues of $\bar{\H}$ to $\epsilon$ times the maximum.
\STATE Do the Cholesky decomposition $\hat{\H} = \C \C^T$
\STATE Let $\bar{\V} = \V \C^{-T}$ (do this in-place; $\C^{-T}$ is upper triangular)
\STATE $\a \gets 0 \in \Re^{K+1}$
\STATE Find the optimum $\a^*$ with BFGS for about $K$ iterations using the subset ${\cal C}_n$, with objective function measured at $\btheta + \bar{\V}\a$ and gradient $\bar{\V}^T \g$ (where $\g$ is the gradient w.r.t. $\btheta$).
\STATE $\d_\mathrm{prev} \gets \bar{\V}\a^*$
\STATE $\btheta \gets \btheta + \d_\mathrm{prev}$
\ENDFOR
\end{algorithmic}
\end{algorithm}
The complete algorithm is given as Algorithm~\ref{alg:ksd}. The most important
parameter is $K$, the dimension of the Krylov subspace (e.g. 20). The flooring
constant $\epsilon$ is an unimportant parameter; we used $10^{-4}$. The subset sizes may be important; we recommend that ${\cal A}_n $ should be all of the training data, and ${\cal B}_n$ and ${\cal C}_n$ should each be about $1/K$ of the training data, and disjoint from each other but not from ${\cal A}_n$. This is the subset size
that keeps the computation approximately balanced between the gradient computation, subspace
construction and subspace optimization. Implementations of the BFGS algorithm would typically also have parameters: for instance,
parameters of the line-search algorithm and stopping critiera; however, we expect that
in practice these would not have too much effect on performance because
the algorithm is likely to converge almost exactly (since the subspace dimension and the
number of iterations are about the same).
\section{Experiments}
\label{sec:exp}
To evaluate KSD, we performed several experiments to compare it with SGD and with
other second order optimization methods, namely L-BFGS and HF. We report both
training and cross validation errors, and running time (we terminated the algorithms
with an early stopping rule using held-out validation data). Our
implementations of both KSD and HF are based on Matlab using
Jacket\footnote{www.accelereyes.com} to perform the expensive matrix operations
on a Geforce GTX580 GPU with 1.5GB of memory.
\subsection{Datasets and models}
Here we describe the datasets that we used to compare KSD to other methods.
\begin{table}
\centering
\begin{tabular}{l|c|c|c|c|c|c}
\hline
Dataset&Train smp.&Test smp.&Input&Output&Model&Task\\
\hline
CURVES&20K&10K&784 (bin.)&784 (bin.)&400-200-100-50-25-5&AE\\
MNIST$_{AE}$&60K&10K&784 (bin.)&784 (bin.)&1000-500-250-30&AE\\
MNIST$_{CL}$&60K&10K&784 (bin.)&10 (class)&500-500-2000&Class\\
MNIST$_{CL,PT}\footnotemark[1]$&60K&10K&784 (bin.)&10 (class)&500-500-2000&Class\\
Aurora&1.2M&100K\footnotemark[2]&352 (real)&56 (class)&512-1024-1536&Class\\
Starcraft&900&100&5077 (mix)&8 (class)&10&Class\\
\hline
\end{tabular}
\caption{Datasets and models used in our setup.}
\label{tab:models}
\end{table}
\begin{itemize}
\item CURVES: Artificial dataset consisting of curves at $28\times28$ resolution. The dataset consists of 20K training samples, and 10K testing samples. We considered an autoencoder network, as in \cite{HinSal06}.
\item MNIST: Single digit vision classification task. The digits are $28\times28$ pixels, with a 60K training, and 10K testing samples. We considered both an autoencoder network, and classification \cite{HinSal06}.
\item Aurora: Spoken digits dataset, with different levels of real noise (airport, train station, ...). We used PLP features and performed classification of 56 English phones. These frame level phone error rates are the ones reported in Table~\ref{tab:results}. Also reported in the text are Word Error Rates, which were produced by using the phone posteriors in a Tandem system, concatenated with standard MFCC to train a Hidden Markov Model with Gaussian Mixture Model emissions. Further details on the setup can be found in \cite{VinyalsICASSP11}.
\item Starcraft: The dataset consists of a real time strategy video game sequences from 1000 games. The goal is to predict the strategy the opponent chose based on a fully observed game sequence after five minutes, and features contain orderings between buildings, presence/absence features, or times that certain buildings were built.
\end{itemize}
The models (i.e. network architectures) for each dataset are summarized in Table~\ref{tab:models}. We tried to explore a wide variety of models covering different sizes, input and output characteristics, and tasks. Note that the error reported for the autoencoder (AE) task is the L2 norm squared between input and output, and for the classification (Class) task is the classification error (i.e. 100-accuracy). The non linearities considered were logistic functions for all the hidden layers except for the ``coding'' layer (i.e. middle layer) in the autencoders, which was linear, and the visible layer for classification, which was softmax.
\footnotetext[1]{For MNIST$_{CL,PT}$ we initialize the weights using pretraining RBMs as in \cite{HinSal06}. In the other experiments, we did not find a significant difference between pretraining and random initialization as in \cite{martens2010deep}.}
\footnotetext[2]{We report both classification error rate on a 100K CV set, and word error rate on a 5M testing set with different levels of noise}
\subsection{Results and discussion}
Table~\ref{tab:results} summarizes our results. We observe that KSD converges faster than HF, and tends to lead to lower generalization error. Our implementation for the two methods is almost identical; the steps that dominate the computation (computing objective functions, gradients and Hessian or Gauss-Newton products) are shared between both and are computed on a GPU.
For all the experiments we used the Gauss-Newton matrix (unless otherwise specified). The dimensionality of the Krylov subspace was set to 20, the number of BFGS iterations was set to 30 (although in many cases the optimization on the projected gradients converged before reaching 30), and an L2 regularization term was added to the objective function. However, motivated by the observation that on CURVES, HF tends to use a large number of iterations, we experimented with a larger subspace dimension of $K=80$ and these are the numbers we report in Table~\ref{tab:results}.
For compatibility in memory usage with KSD, we used a moving window of size 10 for the L-BFGS methods.
We do not report SGD performance in Figures~\ref{fig:aurora} and~\ref{fig:curves} as it was worse
than L-BFGS.
When using HF or KSD, pre-training helped significantly in the MNIST classification task, but not for the other tasks (we do not show the results with pre-training in the other cases; there was no significant difference). However, when using SGD or CG for optimization (results not shown), pre-training helped on all tasks except Starcraft (which is not a deep network). This is consistent with the notion put forward in~\cite{martens2010deep} that it might be possible to do away with the need for pre-training if we use powerful second-order optimization methods. The one exception to this, MNIST, has zero training error when using HF and KSD, which is consistent with a regularization interpretation of pre-training. This is opposite to the conclusions reached in~\cite{ErhanAISTATS2009} (their conclusion was that pre-training helps by finding a better ``basin of attraction''), but that paper was not using these types of optimization methods. Our experiments support the notion that when using advanced second-order optimization methods and when overfitting is not a major issue, pre-training is not necessary. We are not giving this issue the attention it deserves, since the primary focus of this paper is on our optimization method; we may try to support these conclusions more convincingly in future work.
\begin{table}
\centering
\begin{tabular}{l|c|c|c|c|c|c}
\hline
&\multicolumn{3}{|c|}{HF}&\multicolumn{3}{|c}{KSD}\\
\hline
Dataset&Tr. err.&CV err.&Time&Tr. err.&CV err.&Time\\
\hline
CURVES&0.13& \textbf{0.19}&1 &0.17 &0.25 &0.2 \\
MNIST$_{AE}$&1.7& 2.7&1 &1.8 &\textbf{2.5} &0.2 \\
MNIST$_{CL}$&0\%& 2.01\%&1 &0\% &\textbf{1.70\%} &0.6 \\
MNIST$_{CL,PT}$&0\%& 1.40\%&1 &0\% &\textbf{1.29\%} &0.6 \\
Aurora&5.1\%& 8.7\%&1 &4.5\% &\textbf{8.1\%} &0.3 \\
Starcraft&0\%& 11\%&1 &0\% &\textbf{5\%} &0.7 \\
\hline
\end{tabular}
\caption{Results comparing two second order methods: Hessian Free and Krylov Subspace Descent. Time reported is relative to the running time of HF (lower than 1 means faster).}
\label{tab:results}
\end{table}
In Figures~\ref{fig:aurora}~and~\ref{fig:curves}, we show the convergence of KSD and HF with both the Hessian and Gauss-Newton matrices. HF eventually ``gets stuck'' when using the Hessian; the algorithm was not designed to be used for non-positive definite matrices. Even before getting stuck, it is clear that it does not work well with the actual Hessian. Our method also works better with the Gauss-Newton matrix than with the Hessian, although the difference is smaller. Our method is always faster than HF and L-BFGS.
\begin{figure}[ht]
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=\linewidth]{aurora2.eps}
\caption{Aurora convergence curves for various algorithms.}
\label{fig:aurora}
\end{minipage}
\hspace{0.5cm}
\begin{minipage}[b]{0.5\linewidth}
\centering
\includegraphics[width=\linewidth]{curves2.eps}
\caption{CURVES convergence curves for various algorithms.}
\label{fig:curves}
\end{minipage}
\end{figure}
\section{Conclusion and future work}
In this paper, we proposed a new second order optimization method. Our approach relies on efficiently computing the matrix-vector product between the Hessian (or a PSD approximation to it), and a vector. Unlike Hessian Free (HF) optimization, we do not require the approximation of the Hessian to be PSD, and our method requires fewer heuristics; however, it requires more memory.
Our planned future work in this direction includes investigating the circumstances under which pre-training is necessary: that is, we would like to confirm our statement that pre-training is not necessary when using sufficiently advanced optimization methods, as long as overfitting is not the main issue. Current work shows that the presented method is also able to efficiently train recursive neural networks, with no need to use the structural damping of the Gauss-Newton matrix proposed in~\cite{MartensRNN}.
\bibliographystyle{plain}
|
\section{Introduction}
In this paper we continue the study of the classical integrability of Green--Schwarz superstrings in $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ superbackgrounds, whose spectrum contains non--supercoset worldsheet degrees of freedom, initiated in \cite{Sorokin:2010wn,Sorokin:2011rr}.
The integrability properties of superstrings on semi--symmetric coset superspaces $G/H$ with $\mathbb Z_4$--grading are, by now, very well understood. The prescription for constructing a Lax representation of the equations of motion of $2d$ sigma--models on the supercoset $G/H$ (that generates an infinite set of conserved charges) has been proposed in \cite{Bena:2003wd} and applied to various concrete examples \cite{Adam:2007ws} including the maximally supersymmetric type IIB $AdS_5\times S^5$ superstring whose target superspace is $\frac{PSU(2,2|4)}{SO(1,4)\times SO(5)}$ and an $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$ sigma--model \cite{Arutyunov:2008if,Stefanski:2008ik} which is a kappa--symmetry gauge--fixed sub--sector of the Green--Schwarz superstring on a type IIA $AdS_4\times CP^3$ superspace \cite{Gomis:2008jt}. Other examples of interest, in particular in the AdS/CFT context, are string sigma--models on $\frac{PSU(1,1|2)\times PSU(1,1|2)}{SU(1,1)\times SU(2)}$ whose bosonic body is the $6d$ symmetric space $AdS_3\times S^3$ and on $\frac{D(2,1;\alpha)\times D(2,1;\alpha)}{SO(1,2)\times SO(3)\times SO(3)}$ having $AdS_3\times S^3\times S^3$ as its bosonic subspace \cite{Babichenko:2009dk}.
These cases are related to $10d$ superstrings compactified on $AdS_3\times S^3\times M_4$ (where $M_4$ is $T^4$ or $S^3\times S^1$) that preserve 16 target--space supersymmetries.
Another example, which we will consider here, is a superstring on the coset superspace $\frac{PSU(1,1|2)}{SO(1,1)\times U(1)}$ with the $4d$ bosonic subspace $AdS_2\times S^2$ and eight supersymmetric Grassmann--odd directions. This model is a consistent truncation of a $10d$ Green--Schwarz superstring on $AdS_2\times S^2\times T^6$ or $AdS_2\times S^2\times CY^3$ (see \cite{Sorokin:2011rr} for more details and references). It is useful to have a supercoset description which captures the full 10d bosonic geometry of $AdS_2\times S^2\times T^6$ rather than a truncation to 4d. This can be achieved by noting that $AdS_2\times S^2\times\mathbb R^6$, with eight fermionic directions, is described by the supercoset $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$, where the semi-direct product with $E(6)$, the Euclidean group in six dimensions, accounts for the $\mathbb R^6$ factor. Since $AdS_2\times S^2\times T^6$ is locally the same as $AdS_2\times S^2\times\mathbb R^6$, and we will only be interested in the local geometry, this gives us a (local) supercoset description of $AdS_2\times S^2\times T^6$.
Among the above examples only the $AdS_5\times S^5$ superstring is maximally supersymmetric in the $10d$ target space. Its number of supersymmetries and corresponding string fermionic modes is 32 coinciding with the number of Grassmann--odd directions of $\frac{PSU(2,2|4)}{SO(1,4)\times SO(5)}$. In other words, all the worldsheet fermionic modes of the $AdS_5\times S^5$ string are in one to one correspondence with the Grassmann directions of the supercoset space which fully describes the supergeometry of the type IIB $AdS_5\times S^5$ supergravity solution. As a consequence, the prescription of \cite{Bena:2003wd} for the construction of a zero--curvature Lax connection from the $\mathbb Z_4$--graded components of the Cartan form on $\frac{PSU(2,2|4)}{SO(1,4)\times SO(5)}$ demonstrates the classical integrability of the full Green--Schwarz superstring in the $AdS_5\times S^5$ superbackground which coincides with $\frac{PSU(2,2|4)}{SO(1,4)\times SO(5)}$.
Other, less supersymmetric, cases turn out to be more involved. For instance, the Green--Schwarz superstring on $AdS_4\times CP^3$ is invariant under 24 target--space supersymmetries that generate the superisometry group $OSp(6|4)$. The type IIA superspace, in which the string moves, has 32 fermionic directions while the supercoset $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$ only has 24. This means that only 24 of the 32 fermionic modes on the string worldsheet can be associated with the supercoset Grassmann--odd directions, while the 8 remaining fermionic modes (corresponding to broken target--space supersymmetries) do not have this group--theoretical meaning. In fact, the complete type IIA $AdS_4\times CP^3$ superspace is not a supercoset, though it has the $OSp(6|4)$ isometries. Its geometry is much more complicated and reduces to that of $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$ only in the sub--superspace in which the 8 non--supersymmetric fermionic coordinates are put to zero \cite{Gomis:2008jt}. In the Green--Schwarz superstring sigma--model on $AdS_4\times CP^3$ superspace these eight non--supercoset fermionic modes can be put to zero by partially gauge fixing the kappa--symmetry for almost all classical configurations of the string. This however is not possible when the string motion is restricted to the $AdS_4$ subspace \cite{Arutyunov:2008if,Gomis:2008jt} or when the string forms a worldsheet instanton by wrapping a $CP^1$ cycle in $CP^3$ \cite{Cagnazzo:2009zh}. In these cases the supercoset kappa--symmetry gauge is inadmissible, and the non--coset fermions carry physical worldsheet degrees of freedom.\footnote{Subtleties of gauge fixing kappa-symmetry in a way consistent with the light--cone gauge in a near plane--wave limit of $AdS_4 \times CP^3$ has been discussed in \cite{Astolfi:2009qh}. } As a result, the construction of a Lax connection of the Green--Schwarz superstring in the full $AdS_4\times CP^3$ superspace, in general, should include the contribution of the non--coset fermions which will thus modify the form of the supercoset Lax connection of \cite{Arutyunov:2008if,Stefanski:2008ik} by terms whose structure is not captured by the prescription of \cite{Bena:2003wd}.
The situation becomes even more interesting and complicated in less supersymmetric cases such as strings on $AdS_3\times S^3\times M_4$ and $AdS_2\times S^2\times T^6$. For instance, as we have already mentioned, in $AdS_2\times S^2\times T^6$ only 8 target--space supersymmetries corresponding to the Grassmann--odd directions of $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$ are preserved and hence the other 24 fermionic modes of the Green--Schwarz superstring cannot be associated with the supercoset. Moreover, since there are only 16 kappa--symmetries, they can gauge away not more than 16 of these fermions, so that at least 8 of the non--coset worldsheet fermions carry physical degrees of freedom and will always contribute to the structure of the Lax connection of the complete $10d$ theory.
To deal with the non--coset fermions, an alternative prescription for constructing Lax connections has been proposed in \cite{Sorokin:2010wn}. It uses the Noether currents of the isometries of the (super)background as building blocks of the Lax connection and can thus be applied to more general cases than the $G/H$ sigma--models with $\mathbb Z_4$--grading. Using this procedure, zero--curvature Lax connections for superstrings on $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ have been constructed up to second order in the 32 fermionic modes, respectively, in \cite{Sorokin:2010wn} and \cite{Sorokin:2011rr}. In addition, in \cite{Sorokin:2010wn} a Lax connection to all orders in non--coset fermions has been constructed in a special kappa--symmetry gauge of \cite{Grassi:2009yj} in the sub--sector of the $AdS_4\times CP^3$ superstring which cannot be reduced to the $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$ supercoset, thus providing evidence for the classical integrability of the complete theory.
When the non--coset fermions are put to zero, the Lax connections of \cite{Sorokin:2010wn,Sorokin:2011rr} are related to those of
\cite{Bena:2003wd,Adam:2007ws,Arutyunov:2008if,Stefanski:2008ik} (truncated to the second order in the \emph{coset} fermions) by a superisometry gauge transformation that depends on the spectral parameter \cite{Sorokin:2010wn}. To understand how the presence of the \emph{non--coset} fermions modifies e.g. the algebraic curve constructed with the use of the $\mathbb Z_4$--graded supercoset Lax connection and, hopefully, to reveal a role of the non--coset fermionic and bosonic modes in the corresponding Bethe--ansatz techniques, it seems useful to have at hand an explicit expression which demonstrates how the $\mathbb Z_4$--graded supercoset Lax connection gets generalized by terms depending on the non--coset fermions. In this paper we provide such an expression for the Lax connections of the superstring on $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ to all orders in the coset fermions and up to the second order in the non--coset fermions. Interestingly enough, the Lax connections of the $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ superstrings have formally a very similar form. This is because their numbers of target--space supersymmetries complement each other to $32=24+8$, the maximal number of \emph{10d} type II supersymmetries, and the projectors which split 32--component fermions into 24-- and 8--component ones are the same in both of the cases.
This similarity actually has helped us to guess the form of the $AdS_2\times S^2\times T^6$ Lax connection upon having constructed the $AdS_4\times CP^3$ one using the knowledge of the complete $AdS_4\times CP^3$ supergeometry and the superstring equations of motion. As a byproduct, this has also allowed us to get corrections due to the non--coset fermions to the geometry of the $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$ supercoset thus obtaining the form of the geometry of the complete type IIA $AdS_2\times S^2\times T^6$ superspace to all orders in the \emph{coset} fermions and to the second order in the \emph{non--coset} ones. These results make explicit the general discussion of \cite{Sorokin:2011rr} about the structure of the $AdS_2\times S^2\times T^6$ supergeometry which ensures that the $\frac{PSU(1,1|2)}{SO(1,1)\times U(1)}$ sigma--model is a consistent truncation of the complete $10d$ superstring action on $AdS_2\times S^2\times T^6$.
An important property of the obtained form of the Lax connection is that it is invariant under the $\mathbb Z_4$ transformations of the superisometry generators provided that the spectral parameter \textbf{x} gets replaced with its inverse $\frac{1}{\mathbf x}$. This demonstrates that the contribution of the non--coset fermions does not spoil the $\mathbb Z_4$--symmetry of the \emph{supercoset} Lax connection which is of crucial importance for the Bethe ansatz equations, both classical and quantum \cite{Beisert:2005bm,Gromov:2008bz,Babichenko:2009dk,SchaferNameki:2010jy}.
The paper is organized as follows.
In Section 2 we explain our conventions and notation and describe some general properties of the supercoset Lax connection, and the relation of its zero curvature condition to the equations of motion of the corresponding $2d$ dynamical system.
In Section 3 we extend the supercoset Lax connection with contributions coming from the string fermionic modes associated with broken targetspace supersymmetries. In Section \ref{AdS4} we sketch the construction of the Lax connection of the $AdS_4\times CP^3$ superstring starting from that of the $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$ sigma--model and modifying it with terms containing the non--coset fermions (which we call $\upsilon$) in such a way that the zero--curvature condition is satisfied order by order in $\upsilon$ if the worldsheet fields obey the superstring equations of motion.
In Section \ref{AdS2} we pass to the consideration of the $AdS_2\times S^2\times T^6$ case and assume that its Lax connection has a similar form to that of the $AdS_4\times CP^3$ superstring but with the role of the coset and non--coset fermions interchanged and with an appropriate redefinition of the form of the gamma--matrices involved in the construction. We then require that this Lax connection has zero curvature, derive from this condition the equations of motion of the $AdS_2\times S^2\times T^6$ superstring and reconstruct the geometry of its target superspace. Namely, we find the bosonic and fermionic vielbeins, the spin connection and the NS--NS three--form superfield strength of the $AdS_2\times S^2\times T^6$ superspace to all orders in the fermions parametrizing the supercoset $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$ and to the second order in the non--coset fermions.
In Section 4 we demonstrate that the obtained Lax connections are invariant under the $\mathbb Z_4$--transformations and discuss their relation to conserved currents and the Lax connections constructed in \cite{Sorokin:2010wn,Sorokin:2011rr}.
In the Conclusions we discuss open problems, possible generalizations and applications of the results obtained.
\section{Setting the stage}
\subsection{Main notation and conventions}
\label{sec:notation}
We use the metric with the `mostly plus' signature $(-,+,\cdots,+)$.
Generically, the tangent space vector indices are labelled by letters from the beginning of the
Latin alphabet, while letters from the middle of the Latin alphabet stand for curved (world)
indices. The spinor indices are labelled by Greek letters from the beginning of the alphabet, while their curved (world) counterparts are denoted by letters from the middle of the Greek alphabet.
The bosonic coordinates of the ten--dimensional type IIA target superspace in which the string moves are denoted by $X^M$ $(M=0,1,\cdots,9)$ and the Grassmann--odd coordinates are denoted by $\Theta^{\mu}$ ($\mu=1,\cdots, 32)$. Since we consider the string sigma--model we shall always assume that $X^M$ and $\Theta^{\mu}$ depend on the string worldsheet variables $\xi^i=(\tau,\sigma)$.
The geometry of the target superspace is encoded in the form of the vector $\mathcal E^A(X,\Theta)$ and spinor $\mathcal E^\alpha(X,\Theta)$ supervielbeins, and spin connection $\Omega^{AB}(X,\Theta)$. In the string sigma--model these one--forms are pulled back on the string worldsheet, which will always be implicit in what follows, \emph{e.g.} $\mathcal E^A(X,\Theta)=d\xi^i(\partial_iX^M\mathcal E_M{}^A(X,\Theta)+\partial_i\Theta^\mu\mathcal E_\mu{}^A(X,\Theta))$. The $10d$ supergeometry is subject to the basic torsion constraint which we choose to be
\begin{eqnarray}\label{TA}
T^A\equiv d\mathcal E^A+\mathcal E^B\Omega_B{}^A=-i\mathcal E\Gamma^A\mathcal E+i\mathcal E^A\,\mathcal E\lambda+\frac{1}{3}\mathcal E^A\,\mathcal E^B\,\partial_B\,\phi
\end{eqnarray}
and the NS--NS three--form superfield strength is constrained as in \cite{Grassi:2009yj}
\begin{equation}
H=
-i\mathcal E^A\,\mathcal E\Gamma_A\Gamma_{11}\mathcal E
+i\mathcal E^B\mathcal E^A\,\mathcal E\Gamma_{AB}\Gamma_{11}\lambda
+\frac{1}{3!}\mathcal E^C\mathcal E^B\mathcal E^A\,H_{ABC}\,,
\label{eq:H}
\end{equation}
where $\lambda_\alpha(X,\Theta)$ is the dilatino superfield, $\phi(X,\Theta)$ is the dilaton and $\partial_A=\mathcal E_A{}^M\partial_M+\mathcal E_A{}^\mu\partial_\mu$. The dilatino superfield is not independent but is proportional to the spinor derivative of the dilaton \cite{Howe:2004ib}
\begin{equation}\label{lambda}
\lambda_\alpha=-\frac{i}{3}\,\partial_\alpha\,\phi\, :=-\frac{i}{3}\left(\mathcal E_\alpha{}^M\partial_M\phi+\mathcal E_\alpha{}^\mu\partial_\mu\phi\right)\,.
\end{equation}
The matrices $\mathcal E_A{}^M$, $\mathcal E_A{}^\mu$, etc. in the definition of $\partial_A$ and $\partial_\alpha$ are the inverse supervielbeins.
We shall consider classical superstrings in $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ superbackgrounds of type IIA supergravity. As we will see many expressions turn out to be very similar which allows us to treat both cases simultaneously.
Nevertheless, strings in the $AdS_2\times S^2\times T^6$ superbackground of type IIB supergravity can also be treated in a similar fashion with only slight modifications of our formulas due to the same chirality of the $10d$ Majorana--Weyl spinors (see \cite{Sorokin:2011rr}).
\\{}\\
$\mathbf{AdS_4\times CP^3}$ is
parametrized by the $AdS_4$ coordinates $x^m$ $(m=0,1,2,3)$ and the $CP^3$ coordinates $y^{m'}$ $(m'=4,5,6,7,8,9)$. The vielbeins along $AdS_4$ are $e^a(x)=dx^m\,e_m{}^{a}(x)$ ($a=0,1,2,3$) and along $CP^3$ are $e^{a'}(y)=dy^{m'}e_{m'}{}^{a'}(y)$. The $10d$ vielbein is then $e^A(X^M)=(e^a(x),e^{a'}(y))$.
The $AdS_4$ curvature is
\begin{equation}
R_{ab}{}^{cd}=\frac{8}{R^2}\,\delta^c_{[a}\,\delta_{b]}^d \,,\qquad R^{ab}=-\frac{4}{R^2}\,e^a\,e^b\,,
\end{equation}
where $R$ is the $CP^3$ radius or twice the $AdS_4$ radius, and the $CP^3$ curvature is
\begin{equation}
R_{a'b'}{}^{c'd'}=-\frac{2}{R^2}\,(\delta^{c'}_{[a'}\,\delta_{b']}^{d'}+J_{[a'}{}^{c'}\,J_{b']}{}^{d'}+J_{a'b'}J^{c'd'})\,,
\end{equation}
where $J^{a'b'}$ is the K\"ahler form on $CP^3$.
\\{}\\
$\mathbf{AdS_2\times S^2\times T^6}$ is parametrized by the $AdS_2$ coordinates $x^m$ $(m=0,1)$, the $S^2$ coordinates $x^{\hat m}$ $(\hat m=2,3)$ and those of $T^6$ $y^{m'}$ $(m'=4,5,6,7,8,9)$. The corresponding vielbeins are $e^{ a}=dx^{m}\,e_{m}{}^{a}(x)$ (${a}=0,1$),
$e^{\hat a}=dx^{\hat
m}\,e_{\hat m}{}^{\hat a}(\hat x)$ (${\hat a}=2,3$) and $e^{a'}(y)=dy^{a'}$. We will often combine the $AdS_2$ and $S^2$ indices into $\underline a=(a,\hat a)=0,1,2,3$.
The $AdS_2$ curvature is
\begin{equation}
R_{ab}{}^{cd}=\frac{8}{R^2}\,\delta^c_{[a}\,\delta_{b]}^d\,,\qquad R^{ab}=-\frac{4}{R^2}\,e^a\,e^b\,,
\end{equation}
where $R$ is \emph{twice} the $AdS_2$ (or $S^2$) radius, and the $S^2$ curvature is
\begin{equation}
R_{\hat a\hat b}{}^{\hat c\hat d}=-\frac{8}{R^2}\,\delta^{\hat c}_{[\hat a}\,\delta_{\hat b]}^{\hat d}\,,\qquad R^{\hat a\hat b}=\frac{4}{R^2}\,e^{\hat a}\,e^{\hat b}\,.
\end{equation}
The $10d$ curvature of $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ is denoted by $R_{AB}{}^{CD}$.
The $D=10$ gamma--matrices satisfy
\begin{equation}
\{\Gamma^A,\,\Gamma^B\}=2\eta^{AB}\,,\qquad \Gamma^A=(\Gamma^a,\,\Gamma^{a'})\,,\qquad a=0,1,2,3\qquad a'=4,\cdots,9\,.
\end{equation}
We also define
\begin{eqnarray}\label{gammas}
\gamma_5&=&i\Gamma^{0123},\nonumber\\
\gamma_7&=&i\Gamma^{456789},\nonumber\\
\Gamma_{11}&=&\gamma_5\gamma_7\,,
\end{eqnarray}
all of which square to one. The charge conjugation matrix is denoted $\mathcal C$. The matrices $\mathcal C$, $\mathcal C\Gamma_{\hat A\hat B\hat C}$ and $\mathcal C\Gamma_{\hat A\hat B\hat C\hat D}$ are anti-symmetric while $\mathcal C\Gamma_{\hat A}$, $\mathcal C\Gamma_{\hat A\hat B}$ and $\mathcal C\Gamma_{\hat A\hat B\hat C\hat D\hat E}$ are symmetric, where the indices are eleven dimensional, $\hat A=(A,11)$.
Finally we introduce a spinor projection matrix $\mathcal P_8$ which singles out an 8--dimensional subspace of the 32--dimensional space of spinors
\begin{equation}\label{P8}
\mathcal P_8=\frac{1}{8}(2-iJ_{a'b'}\Gamma^{a'b'}\gamma^7)\,,
\end{equation}
where $J_{a'b'}$ is the K\"ahler form on $CP^3$ or $T^6$. The complementary projection matrix which singles out a 24--dimensional subspace is then
\begin{equation}\label{P24}
\mathcal P_{24}=1-\mathcal P_8=\frac{1}{8}(6+iJ_{a'b'}\Gamma^{a'b'}\gamma^7)\,.
\end{equation}
Some useful identities satisfied by the gamma matrices and these projectors are given in Appendix \ref{sec:gammamatrices}.
In the $AdS_4\times CP^3$ case $\mathcal P_{24}$ singles out from $\Theta^\alpha$ 24 fermionic coordinates
$$\vartheta=\mathcal P_{24}\Theta$$
corresponding to the unbroken supersymmetries and, hence, to the Grassmann--odd directions of the supercoset $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$, while the remaining eight
$$\upsilon=\mathcal P_8\Theta$$
are non--supercoset fermions.
In $AdS_2\times S^2\times T^6$ the role of the two projectors gets exchanged. $\mathcal P_{8}$ singles out from $\Theta^\alpha$ 8 fermionic coordinates
$$\vartheta=\mathcal P_{8}\Theta$$
corresponding to the unbroken supersymmetries and, hence, to the Grassmann--odd directions of the supercoset $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$, while the remaining twenty four
$$\upsilon=\mathcal P_{24}\Theta$$
are non--supercoset fermions.
To treat the two cases simultaneously we shall always denote the coset fermions by $\vartheta$ and the non--coset ones by $\upsilon$.
The projector which singles out the \emph{coset} fermions will be denoted by $\mathcal P$, namely
\begin{equation}\label{P}
\vartheta=\mathcal P\Theta, \qquad\upsilon=(1-\mathcal P)\Theta.
\end{equation}
Let us note that as solutions of the type IIA supergravity equations of motion the $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ backgrounds also contain non--zero constant Ramond--Ramond $F_2$ and $F_4$ fluxes which are implicitly encoded in the form of the projectors \eqref{P8} and \eqref{P24}, namely \begin{equation}\label{slash}
\slashed F=-\frac{8i}{R}\mathcal P\gamma_\star,
\end{equation}
where
$$
\slashed F=e^\phi\left(-\frac{1}{2}\Gamma^{AB}\Gamma_{11}F_{AB}+\frac{1}{4!}\Gamma^{ABCD}F_{ABCD}\right)
$$
and $\gamma_\star$ stands for $\gamma_5$ in the $AdS_4\times CP^3$ case and for $\Gamma^{01}\gamma^7$ in the $AdS_2\times S^2\times T^6$ case. The explicit form of the RR fluxes can be found e.g. in \cite{Gomis:2008jt} and \cite{Sorokin:2011rr}\footnote{To be precise, eq. \eqref{slash} holds for the $AdS_2\times S^2\times T^6$ background with a non--zero $F_4$ and $F_2$ flux such that the latter has support on $S^2$ as in eq. (3.3) of \cite{Sorokin:2011rr}.}.
\subsection{Superstring action and equations of motion}
The Green--Schwarz superstring action in a general supergravity background \cite{Grisaru:1985fv}, written in terms of worldsheet differential forms, is
\begin{equation}
S=-\frac{T}{2}\int_\Sigma\,*\mathcal E^A\mathcal E^B\eta_{AB}+T\int_\Sigma\,B\,,
\end{equation}
where the pull--back to the worldsheet of the target--superspace quantities is understood, the star $*$ denotes the Hodge dual operation on the worldsheet and the wedge product of differential forms is implicit.
From this action one gets the superstring equations of motion which have the following form for our choice of the superspace constraints, eq. \eqref{TA} (see also \cite{Gomis:2008jt,Grassi:2009yj}). The fermionic field equations are
\begin{eqnarray}\label{Psi}
\Psi_\alpha\equiv i* {\mathcal E}^A\,(\Gamma_A {\mathcal E})_\alpha
-i {\mathcal E}^A\,(\Gamma_A\Gamma_{11} {\mathcal E})_\alpha
+\frac{i}{2}* {\mathcal E}^A {\mathcal E}_A\,\lambda_\alpha
+\frac{i}{2} {\mathcal E}^A {\mathcal E}^B\,(\Gamma_{AB}\Gamma_{11}\lambda)_\alpha=0\,,
\end{eqnarray}
and the bosonic field equations are
\begin{eqnarray}\label{B}
\mathcal B^A&\equiv&d*\mathcal E^A+*\mathcal E^B\Omega_B{}^A
+i* {\mathcal E}^A\, {\mathcal E}\lambda
+\frac{1}{3}(* {\mathcal E}^A {\mathcal E}^B\, \partial_B\phi-* {\mathcal E}^B {\mathcal E}_B\, \partial^A\phi)
\nonumber\\
&&{}
-i {\mathcal E}\Gamma^A\Gamma_{11} {\mathcal E}-2i {\mathcal E}^B\, {\mathcal E}\Gamma^A{}_B\Gamma_{11}\lambda
+\frac{1}{2} {\mathcal E}^C {\mathcal E}^B\,H^A{}_{BC}=0\,.
\end{eqnarray}
Note that in the $AdS_4\times CP^3$ case $\partial_A\phi=0$ \footnote{To check that in the $AdS_4\times CP^3$ case $\partial_A\phi=0$ one can use the fact that $\phi(\upsilon)$ and $\lambda_\alpha=-\frac{i}{3}\partial_\alpha\phi$ do not depend on $X^M$ and $\vartheta$ (see \cite{Gomis:2008jt}), so we have
$$
0=\partial_M\phi=\mathcal E_M{}^A\partial_A\phi+\mathcal E_M{}^\alpha\partial_\alpha\phi=\mathcal E_M{}^A\partial_A\phi+3i\mathcal E_M{}^{\alpha}\lambda_\alpha=0.
$$
If $\partial_A\phi=0$, from the above equation it follows that the contraction of the gravitino superfield $\mathcal E_M{}^{\alpha}$ with the dilatino $\lambda_\alpha$ of this supergravity solution is zero,
$
\mathcal E_M{}^{\alpha}\lambda_\alpha=0.
$
This can be checked using the explicit expressions for $\mathcal E^{\alpha}$ and $\lambda_\alpha$ derived in \cite{Gomis:2008jt}.}. We shall see that it is also zero in the $AdS_2\times S^2\times T^6$ case to the second order in $\upsilon$ and guess that it may also be true to all orders, because of the similarity between the two cases.
If we put the non--supercoset fermions $\upsilon$ to zero the equations of motion \eqref{Psi} and \eqref{B} reduce to those of the sigma--models on $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$ and $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$, respectively, with the equations of motion of the $T^6$ coordinates decoupled in the latter case. In our conventions the supercoset equations of motion have the following form
\begin{eqnarray}\label{Psicoset}
\Psi_{0\alpha}&=&i*E^A\,(\Gamma_AE)_\alpha-iE^A\,(\Gamma_A\Gamma_{11}E)_\alpha=0\,,\\
\mathcal B^A_0&=&\nabla*E^A-iE\Gamma^A\Gamma_{11}E=0\,,\label{Bcoset}
\end{eqnarray}
where $\nabla*E^A=d*E^A+*E_B\Omega_0^{BA}$ and $\Omega^{AB}_0(X,\vartheta)$ is the supercoset spin connection $\Omega_0^{AB}=\Omega^{AB}|_{\upsilon=0}$, and $E^A(X,\vartheta)=c^{-2}\mathcal E^A|_{\upsilon=0}$ and $E^\alpha(X,\vartheta)=c^{-1}\mathcal E^\alpha|_{\upsilon=0}$ (with $c$ being a constant dilaton factor) are the supercoset supervielbeins. They are the components of the Cartan form valued in the isometry supergroup $G$ (\emph{i.e.} $OSp(6|4)$ or $PSU(1,1|2)\rtimes E(6)$)
\begin{equation}\label{K}
K=g^{-1}dg(X,\vartheta)=\frac{1}{2}\Omega_0^{AB}M_{AB}+E^AP_A+Q_\alpha E^\alpha\,, \qquad g(X,\vartheta) \in G/H\,.
\end{equation}
The algebra of the isometry generators and the explicit form of the supercoset bosonic and fermionic supervielbeins, and the spin connection are given in Appendix A. The isometry algebra is invariant under the following action of the $\mathbb Z_4$--automorphism on the generators $T=(M,P,Q)$
\begin{equation}\label{Z4auto}
\Omega(T)\equiv \Omega^{-1}T\Omega\,\qquad \Omega (M_{AB})=M_{AB},
\qquad
\Omega (P_{A})=-P_{A},
\qquad
\Omega (Q)=-iQ\Gamma _{11}, \qquad \Omega^4=1\,.
\end{equation}
Note that in the $AdS_2\times S^2\times T^6$ case the $T^6$ translation generators $P_{a'}$ also have $\mathbb Z_4$--grading one, as those of $AdS_2\times S^2$.
When the non--coset fermions are non--zero, the supercoset field equations \eqref{Psicoset} and \eqref{Bcoset} acquire non--zero right--hand sides
\begin{equation}\label{PsiB}
\Psi_{0}=\mathcal O(\upsilon)\,,\qquad \mathcal B_0=\mathcal O(\upsilon)
\end{equation}
which should be taken into account when extending the supercoset Lax connection to a zero--curvature Lax connection of the complete theory (see Section 3).
\subsection{Supercoset Lax connection}
The equations of motion of the superstring on the semi--symmetric supercoset spaces with $\mathbb Z_4$--grading \eqref{K} and \eqref{Z4auto} admit a Lax representation which implies classical integrability of the corresponding sigma--model \cite{Bena:2003wd,Adam:2007ws,Arutyunov:2008if,Stefanski:2008ik}. This means that from the components of the Cartan form \eqref{K} pulled--back on the worldsheet and their $2d$ Hodge duals one can construct a Lax connection (depending on a spectral parameter) which has zero curvature provided that the equations of motion \eqref{Psicoset} and \eqref{Bcoset} are satisfied. And vice versa the zero curvature of the Lax connection implies the field equations. In our notation and conventions the supercoset Lax connection has the following form
\begin{equation}\label{Lcoset}
L_{coset}=\frac{1}{2}\Omega_0^{AB}M_{AB}+(1+\alpha_1)E^AP_A+\alpha_2\ast E^AP_A+Q(\beta_2+\beta_1\Gamma_{11})E\,,
\end{equation}
where $\alpha_1$, $\alpha_2$, $\beta_1$ and $\beta_2$ are numerical parameters whose values are determined by requiring the zero--curvature condition
\begin{equation}\label{R0}
dL_{coset}-L_{coset}\wedge L_{coset}=0
\end{equation}
to hold on the mass--shell \eqref{Psicoset} and \eqref{Bcoset}. This gives the following relations between the parameters
\begin{equation}\label{alpha}
\alpha_2^2=2\alpha_1+\alpha_1^2\,
\end{equation}
and
\begin{eqnarray}\label{beta}
\beta_1=\mp\sqrt{\frac{\alpha_1}{2}}\,
\qquad \beta_2=\pm\frac{\alpha_2}{\sqrt{2\alpha_1}
\,.
\end{eqnarray}
They can therefore be expressed in terms of a single spectral parameter $\tt x$ as follows
\begin{equation}\label{x}
\alpha_1=\frac{2{\tt x}^2}{1-{\tt x}^2}, \qquad \alpha_2=\frac{2{\tt x}}{1-{\tt x}^2}\,\qquad \beta_1=-\frac{i{\tt x}}{\sqrt{{\tt x}^2-1}}\,,\qquad \beta_2=\frac{i}{\sqrt{{\tt x}^2-1}}\,.
\end{equation}
It is very useful for further analysis to specify the properties of the $32\times32$ matrix
\begin{equation}\label{eq:V}
V=\beta_2+\beta_1\Gamma_{11}\,,
\end{equation}
which enters the Lax connection \eqref{Lcoset}. It is easily seen to satisfy the relations
\begin{equation}\label{G}
V^2=1+\alpha_1-\alpha_2\Gamma_{11}\,, \quad {VV}^\dagger=\beta_2^2-\beta_1^2=1\,,\quad (V^\dagger)^\alpha{}_\beta=-(\mathcal {C}V^\mathrm{T}\mathcal C)^\alpha{}_\beta=(\beta_2-\beta_1\Gamma_{11})^\alpha{}_\beta\,,
\end{equation}
where $\mathcal C$ denotes the anti-symmetric charge-conjugation matrix (see Section \ref{sec:notation}). Therefore $V\in Sp(32)$.
It is easy to check that the Lax connection \eqref{Lcoset} is invariant under the $\mathbb Z_4$--transformations of the generators \eqref{Z4auto} accompanied by the inversion of the spectral parameter
\begin{equation}\label{Ox}
\Omega(\tt x) = \frac{1}{\tt x}\,,
\end{equation}
which implies that
\begin{equation}\label{Oalpha}
\alpha_1\rightarrow-\alpha_1-2\,,\qquad \alpha_2\rightarrow -\alpha_2\,,\qquad V\rightarrow i\Gamma_{11}V\,.
\end{equation}
Namely,
$$
\Omega(L_{coset}({\tt x}))=\Omega^{-1}L_{coset}(\frac{1}{\tt x})\,\Omega=L_{coset}({\tt x})\,.
$$
Note that in the $AdS_2\times S^2\times T^6$ case the first term of \eqref{Lcoset} may in general include the $SO(6)$ spin connection on $T^6$ whose curvature is zero. Therefore, it can be gauged away by performing a suitable gauge transformation of $L_{coset}$, and the resulting Lax connection will contain only terms associated with the $U(1)^6$ `translations' $P_{a'}$ along $T^6$ which completely decouple from the $PSU(1,1|2)$ part and, therefore, can be taken with arbitrary coefficients.
The explicit dependence of the supercoset Lax curvature on the left--hand sides of the supercoset field equations \eqref{Psicoset} and \eqref{Bcoset} looks as follows
\begin{equation}\label{R01}
dL_{coset}-L_{coset}\wedge L_{coset}=\alpha_2 (\mathcal B_0^AP_A-\frac{1}{R}QV^\dagger\gamma_\star \Psi_0)\,,
\end{equation}
where again $\gamma_\star $ stands for $\gamma^5$ in the $AdS_4\times CP^3$ case and for $\Gamma^{01}\gamma^7$ in the $AdS_2\times S^2\times T^6$ case (see eqs. \eqref{gammas} and \eqref{gammasA}).
\section{The Lax connection of the complete GS superstring to quadratic order in non--coset fermions $\upsilon$}\label{comleteL}
We are now ready to extend the supercoset Lax connection \eqref{Lcoset} with terms that include contributions from the string fermionic modes $\upsilon$ associated with broken target--space supersymmetries
\begin{equation}
L=L_{coset}(X,\vartheta)+\alpha_2L'(X,\vartheta,\upsilon)\,,
\label{eq:Lax-connection}
\end{equation}
where the factor of $\alpha_2$ in front of $L'$ is due to the same factor on the right--hand side of \eqref{R01}. The correction $L'$ which is aimed at canceling the r.h.s. of \eqref{R01}, turns out to take the same form for both the $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ case and looks as follows
\begin{eqnarray}\label{L'}
L'&=&
-\frac{i}{R}Q\gamma_\star \left[
*(E^A+2i\upsilon\Gamma^AE)\,\Gamma_A V \upsilon
-(E^A+2i\upsilon\Gamma^AE)\,\Gamma_A\Gamma_{11} V \upsilon\right]\nonumber\\
&&
-\frac{i}{R}Q\gamma_\star \left[i(\upsilon\Gamma^A\Gamma_{11}E)\,\Gamma_A V \upsilon
+i(\upsilon\Gamma^AE)\,\Gamma_A\Gamma_{11} V \upsilon
\right]
\nonumber\\
&&{}
+(
2i\upsilon\Gamma^A*E+i\upsilon\Gamma^A*\nabla\upsilon-\frac{2}{R}*E^B\,\upsilon\Gamma^A\mathcal P\gamma_\star \Gamma_B\upsilon
)\,P_A
\nonumber\\
&&{}
+(
2i\upsilon\Gamma^A\Gamma_{11}E+i\upsilon\Gamma^A\Gamma_{11}\nabla\upsilon-\frac{2}{R}E^B\,\upsilon\Gamma^A\Gamma_{11}\mathcal P\gamma_\star \Gamma_B\upsilon
)\,P_A
\nonumber\\
&&{}
+\frac{i}{8}(*E^C\,\upsilon\Gamma_C{}^{DE} V ^2\upsilon-E^C\,\upsilon\Gamma_C{}^{DE}\Gamma_{11} V ^2\upsilon)\,R_{DE}{}^{AB}\,M_{AB}
\,,
\end{eqnarray}
where the matrix $V$ has been introduced in \eqref{eq:V}, and $\gamma_\star $, $\mathcal P$ and $R_{DE}{}^{AB}$ are defined for $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ in Section \ref{sec:notation} and Appendix \ref{sec:supercosets}. The covariant derivative is defined with respect to the supercoset spin connection $\Omega_0^{AB}=\Omega^{AB}|_{\upsilon=0}$, i.e. $\nabla \upsilon=(d-\frac{1}{4}\Omega_0^{AB}\Gamma_{AB})\upsilon$.
To construct $L'(X,\vartheta,\upsilon)$ one had to know the form of the right--hand sides of eqs. \eqref{Psicoset} and \eqref{Bcoset}. For the $AdS_4\times CP^3$ case these can be given to all orders in $\upsilon$ since the explicit form of the geometry of the $AdS_4\times CP^3$ superspace is known \cite{Gomis:2008jt}. Because of technical complications, we have however restricted the construction of the Lax connection to the second order in $\upsilon$ only.
In the $AdS_2\times S^2\times T^6$ case the explicit form of the geometry of the complete target superspace is not know, so our strategy was somewhat opposite to that of the $AdS_4\times CP^3$ case. We have assumed that in the $AdS_2\times S^2\times T^6$ case the Lax connection has a similar form to the $AdS_4\times CP^3$ Lax connection, i.e. eq. \eqref{L'} with the appropriate replacement of the supersymmetry projector $\mathcal P$ and the product of gamma--matrices $\gamma_\star $ appearing in the definition of the isometry superalgebra (see Appendix \ref{sec:supercosets}). Then, requiring that the $AdS_2\times S^2\times T^6$ Lax connection has zero curvature we have reconstructed the superstring equations and the form of the $AdS_2\times S^2\times T^6$ superbackground up to the second order in the non--coset fermions and checked that it indeed satisfies the constraints of type IIA supergravity.
\subsection{$AdS_4\times CP^3$ case}\label{AdS4}
Taking the expressions for the quantities defining the full $AdS_4\times CP^3$ supergeometry \cite{Gomis:2008jt} and expanding them to the second order in $\upsilon$ we get
\begin{eqnarray}\label{geometry}
\mathcal E^A&=&c^2(1-\frac{1}{R}\upsilon\gamma^5\upsilon)(E^A+2i\upsilon\Gamma^AE+i\upsilon\Gamma^AD\upsilon)+\mathcal O(\upsilon^3)\,,\nonumber\\
\mathcal P\mathcal E&=&c(1-\frac{1}{2R}\upsilon\Gamma^b\gamma^5\upsilon\,\Gamma_b+\frac{1}{R}\upsilon\Gamma^b\gamma^7\upsilon\,\Gamma_b\Gamma_{11}+\frac{1}{2R}\upsilon\upsilon\,\gamma_5)E+\mathcal O(\upsilon^3)\,,\nonumber\\
(1-\mathcal P)\mathcal E&=&cD\upsilon+\mathcal O(\upsilon^3)\nonumber\,,\\
\lambda&=&\frac{2i}{cR}\gamma^5\upsilon+\mathcal O(\upsilon^3)\nonumber\,,\\
H_{abc}&=&-\frac{12i}{c^2R^2}\upsilon\Gamma_{abc}\Gamma_{11}\upsilon+\mathcal O(\upsilon^3)\,,
\end{eqnarray}
where $c=e^{\frac{1}{6}\phi_0}=\left(\frac{R}{kl_p}\right)^{1/4}$, $\phi_0$ is the value of the dilaton for the $AdS_4\times CP^3$ supergravity solution, $k$ is the `Chern--Simons' level, $l_p$ is the $11d$ Planck length and
\begin{equation}\label{DV1}
D\upsilon=(\nabla+\frac{i}{R}E^a\,\gamma_5\Gamma_a)\upsilon\,.
\end{equation}
Remember that in the above expressions $E^A(X,\vartheta)$, $E^\alpha(X,\vartheta)$ and $\Omega_0^{AB}(X,\vartheta)$ are the components of the Cartan form of the supercoset $\frac{OSp(6|4)}{SO(1,3)\times U(3)}$ (see Section \ref{sec:notation} and Appendix \ref{SG}). Note also that though in the background under consideration the purely bosonic part of the NS-NS flux $H_{ABC}(X)$ is zero, its superfield extension is non--trivial and depends on the non--coset fermionic coordinates.
We can now insert the expressions \eqref{geometry} into the complete equations of motion \eqref{Psi} and \eqref{B} and thus find the corrections to the supercoset equations \eqref{PsiB}. For this we should also know the form of the spin connection $\Omega^{AB}(X,\vartheta,\upsilon)$ which was not derived in \cite{Gomis:2008jt}. The expression for $\Omega^{AB}(X,\vartheta,\upsilon)$ can be obtained by analyzing the torsion constraint \eqref{TA} and has the following form to the second order in $\upsilon$
\begin{eqnarray}\label{Omega}
\Omega^{AB}&=&\Omega_0^{AB}+\frac{2}{R}\Big(
-\delta^A_{a'}\delta^B_{b'}\upsilon\Gamma^{a'b'}\gamma_5E
+\delta^A_a\delta^B_b\,\upsilon\Gamma^{ab}\gamma_5D\upsilon
+\frac{i}{R}\delta^A_{a'}\delta^B_{b'}\,E^c\,\upsilon\Gamma^{a'b'}{}_c\upsilon
\nonumber\\
&&{}
-\frac{2i}{R}\delta^{[A}_{a'}\delta^{B]}_b\,E^{c'}\,\upsilon\Gamma^{a'b}{}_{c'}\upsilon
\Big)+\mathcal O(\upsilon^3)\,.
\end{eqnarray}
Notice that (due to the last term in \eqref{Omega}) the spin connection takes values in the whole $D=10$ Lorentz algebra $so(1,9)$ rather than in the stability subalgebra $so(1,3)\oplus u(3)$ of the bosonic coset $AdS_4\times CP^3$. This reflects the fact that the complete $AdS_4\times CP^3$ superspace is not a supercoset.
The explicit form of the corrections to the equations of motion \eqref{Psicoset} and \eqref{Bcoset} are given in Appendix \ref{eom}. Thus, the correction $L'$ to the supercoset Lax connection which cancels the contributions of the right--hand sides of these equations to the Lax curvature \eqref{R01} has been found to be eq. \eqref{L'}.
\subsection{$AdS_2\times S^2\times T^6$ Lax connection and supergeometry}\label{AdS2}
We now pass to the consideration of the $AdS_2\times S^2\times T^6$ case. As we have already mentioned, a problem that we meet is that here the explicit form of the corresponding $10d$ superspace with 32 fermionic directions is unknown. What is known is the supercoset structure of its sub--superspace obtained by putting to zero 24 fermionic coordinates $\upsilon$ and, on the other hand, the structure of the complete $AdS_2\times S^2\times T^6$ superspace to the second order in the 32 fermions $\Theta$ (see \cite{Sorokin:2011rr} for more details and references). Our goal is to take a step further and to find the explicit form of the $AdS_2\times S^2\times T^6$ superbackground to all orders in the 8 coset fermions $\vartheta$ and to quadratic order in 24 non--coset fermions $\upsilon$. To this end, we assume that the $AdS_2\times S^2\times T^6$ Lax connection has the form of \eqref{eq:Lax-connection} and \eqref{L'} in which now $L_{coset}$ and $L'$ are constructed with the use of the Cartan forms, the curvature of $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$ and the structure of its isometry superalgebra given in Appendix \ref{sec:supercosets}. Namely, in \eqref{L'} we now take $\mathcal P$ to be the supersymmetry projector $\mathcal P_8$ and $\gamma_\star =\Gamma^{01}\gamma^7$ as in the superalgebra of $PSU(1,1|2)\rtimes E(6)$.
Then, requiring that the curvature of the Lax connection \eqref{eq:Lax-connection} and \eqref{L'} vanishes, we get the form of the superstring equations \eqref{Psi} and \eqref{B} in $AdS_2\times S^2\times T^6$ as a deformation of the $\frac{PSU(1,1|2)\rtimes E(6)}{SO(1,1)\times U(1)\times SO(6)}$ sigma--model field equations to the second order in $\upsilon$ and consequently reconstruct to the same order the form of the supervielbeins, the spin connection and the NS--NS superfield strength of the complete $AdS_2\times S^2\times T^6$ superbackground.
Using the expressions given in Appendix \ref{sec:dL} the curvature of the Lax connection \eqref{eq:Lax-connection} with $L'$ given in eq. \eqref{L'} and valued in the $PSU(1,1|2)\rtimes E(6)$ superalgebra of Appendix \ref{sec:supercosets} can be assembled to have the following form
\begin{eqnarray}
dL-LL&=&
\frac{\alpha_2}{c^3}\Big[
\frac{1}{4}(
\upsilon(V^\dagger)^2\Gamma^{\underline{cd}}\Psi
+\frac{ic}{2}\mathcal B^{\underline a}\,\upsilon\Gamma_{\underline a}{}^{\underline{cd}}V^2\upsilon
)\,R_{\underline{cd}}{}^{\underline{ef}}M_{\underline{ef}}
+c\mathcal B^A\,P_A
\nonumber\\
&&{}
+\frac{2i}{R}\upsilon\gamma_\star \Gamma^A(1-\mathcal P)\Psi\,P_A
+\frac{c}{R}\mathcal B^B\,\upsilon\Gamma^A{}_B\gamma_\star \upsilon\,P_A
+\frac{c}{R}\mathcal B^B\,\upsilon\Gamma^A\mathcal P\Gamma_B\gamma_\star \upsilon\,P_A
\nonumber\\
&&{}
-\frac{1}{R}QV^\dagger\gamma_\star \Psi
-\frac{ic}{R}\mathcal B^A\,QV^\dagger\gamma_\star \Gamma_A\upsilon
-\frac{1}{4R^2}QV^\dagger\gamma_\star \Gamma_{BC}\Psi\,\upsilon\Gamma^{BC}\gamma_\star \upsilon
\nonumber\\
&&{}
-\frac{1}{2R^2}QV^\dagger\gamma_\star \gamma_7\Gamma_B\Psi\,\upsilon\Gamma^B\gamma_\star \gamma_7\upsilon
-\frac{1}{2R^2}QV^\dagger\gamma_\star \Gamma_B\Gamma_{11}\Psi\,\upsilon\Gamma^B\gamma_\star \Gamma_{11}\upsilon
\nonumber\\
&&{}
-\frac{1}{2R^2}QV^\dagger\gamma_\star \Gamma_{\underline ab'}\gamma_5\gamma_\star \Psi\,\upsilon\Gamma^{\underline ab'}\gamma_5\upsilon
-\frac{1}{4R^2}QV^\dagger\gamma_\star \Gamma_{\underline{ab}c'}\gamma_\star \Psi\,\upsilon\Gamma^{\underline{ab}c'}\upsilon
\nonumber\\
&&{}
+\frac{\alpha_2}{2R^2}\Big(
\upsilon\Gamma^{\underline a}\gamma^5\upsilon\,QV\gamma^7\gamma_\star \Gamma_{\underline a}\gamma_\star \Psi
+\upsilon\Gamma^{\underline a}\gamma^7\upsilon\,QV\gamma^5\gamma_\star \Gamma_{\underline a}\gamma_\star \Psi
\nonumber\\
&&{}
+2\upsilon\Gamma^{a'}\gamma_\star \Gamma_{11}\upsilon\,QV\Gamma_{a'}\gamma_\star \Psi
+2\upsilon\Gamma^{a'}\gamma_\star \upsilon\,QV\Gamma^{11}\Gamma_{a'}\gamma_\star \Psi
\Big)
\Big]+\mathcal O(\upsilon^3)\,,
\label{eq:dLAdS2}
\end{eqnarray}
where, if the curvature is zero, $\Psi_\alpha$ and $\mathcal B^A$ should vanish and hence should coincide with the equations of motion of the $AdS_2\times S^2\times T^6$ superstring.
Comparing the form of these $\Psi_\alpha$ and $\mathcal B^A$ with the Green--Schwarz superstring equations \eqref{Psi} and \eqref{B} we find that the $AdS_2\times S^2\times T^6$ supergeometry is described by the following supervielbeins, dilatino and spin connection
\begin{eqnarray}
\mathcal E^A&=&c^2(1+\frac{1}{R}\upsilon\gamma_\star \upsilon)(E^A+2i\upsilon\Gamma^AE+i\upsilon\Gamma^AD\upsilon)
+\mathcal O(\upsilon^3)\nonumber\\
\mathcal P\mathcal E&=&
c\mathcal P(1
+\frac{1}{2R}\upsilon\Gamma^B\gamma_\star \gamma_7\upsilon\,\Gamma_B\gamma_7-\frac{1}{2R}\upsilon\Gamma^B\gamma_\star \Gamma_{11}\upsilon\,\Gamma_B\Gamma_{11}-\frac{1}{4R}\upsilon\Gamma^{BC}\gamma_\star \upsilon\,\Gamma_{BC})E
+\mathcal O(\upsilon^3)\nonumber\\
(1-\mathcal P)\mathcal E&=&c[D\upsilon
+(1-\mathcal P)(\frac{1}{2R}\upsilon\Gamma^B\gamma_\star \gamma_7\upsilon\,\Gamma_B\gamma_7
-\frac{1}{2R}\upsilon\Gamma^B\gamma_\star \Gamma_{11}\upsilon\,\Gamma_B\Gamma_{11}
-\frac{1}{4R}\upsilon\Gamma^{BC}\gamma_\star \upsilon\,\Gamma_{BC}
\nonumber\\
&&{}
-\frac{1}{2R}\upsilon\Gamma^{\underline bc'}\Gamma_{11}\upsilon\,\gamma_\star \Gamma_{\underline bc'}\Gamma_{11}
-\frac{1}{4R}\upsilon\Gamma^{\underline{bc}d'}\upsilon\,\gamma_\star \Gamma_{\underline{bc}d'}
)E]
+\mathcal O(\upsilon^3)\nonumber\\
\lambda&=&-\frac{2i}{cR}\gamma_\star \upsilon+\mathcal O(\upsilon^3)\,,
\label{eq:AdS2vielbeins}
\end{eqnarray}
\begin{eqnarray}
\Omega^{AB}&=&\Omega_0^{AB}
-\frac{2}{R}\delta^{[A}_{a'}\delta^{B]}_{b'}\upsilon\gamma_\star \Gamma^{a'b'}E
-\frac{1}{R}\upsilon\Gamma^{[A}(1-\mathcal P)\Gamma^{B]}\gamma_\star D\upsilon
-\frac{1}{R}\upsilon\gamma_\star \Gamma^{[A}(1-\mathcal P)\Gamma^{B]}D\upsilon
\nonumber\\
&&{}
+\frac{i}{R^2}\delta^{[A}_{a'}\delta^{B]}_{b'}E^C\,\upsilon\Gamma^{a'}(1-\mathcal P)\Gamma^{b'}\gamma_\star \Gamma_C\gamma_\star \upsilon
+\frac{i}{R^2}\delta^{[A}_{a'}\delta^{B]}_{b'}E^{c'}\,\upsilon\Gamma^{a'}\mathcal P\Gamma^{b'c'}\upsilon
\nonumber\\
&&{}
-\frac{2i}{R^2}\delta^{[A}_{a'}\delta^{B]}_{\underline b}E^{c'}\,\upsilon\Gamma^{a'}(1-\mathcal P)\Gamma^{\underline b}\Gamma_{c'}\upsilon
+\frac{2i}{R^2}\delta^{[A}_{a'}\delta^{B]}_bE^{\hat c}\,\upsilon\Gamma^{a'b}{}_{\hat c}\upsilon
\nonumber\\
&&{}
-\frac{2i}{R^2}\delta^{[A}_{a'}\delta^{B]}_{\hat b}E^c\,\upsilon\Gamma^{a'\hat b}{}_c\upsilon
-\frac{2i}{R^2}\delta^{[A}_a\delta^{B]}_{\underline b}E^{c'}\,\upsilon\Gamma^{a\underline b}{}_{c'}\upsilon
+\mathcal O(\upsilon^3)\,,
\label{eq:AdS2Omega}
\end{eqnarray}
where
\begin{eqnarray}\label{Dv}
D\upsilon=(\nabla+\frac{i}{R}E^B\,(1-\mathcal P)\Gamma_B\gamma_\star )\upsilon\,
\end{eqnarray}
and the constant $c=e^{\frac{1}{6}\phi_0}$, where $\phi_0$ is the dilaton vacuum expectation value of the $AdS_2\times S^2\times T^6$ supergravity solution.
Finally, the NS--NS three--form field strength turns out to be
\begin{eqnarray}
H_{ABC}&=&
\frac{6i}{c^2R^2}\Big(
\upsilon\gamma_\star \Gamma_{ABC}\Gamma_{11}\gamma_\star \upsilon
-\delta_{[A}^{\underline a}\,\upsilon\gamma_\star \Gamma_B\mathcal P\Gamma_{C]}\Gamma_{\underline a}\gamma_\star \Gamma_{11}\upsilon
-\delta_{[A}^d\delta_B^e\,\upsilon\Gamma_{C]de}\Gamma_{11}\upsilon
\nonumber\\
&&{}
+\delta_{[A}^{\hat d}\delta_B^{\hat e}\,\upsilon\Gamma_{C]\hat d\hat e}\Gamma_{11}\upsilon
\Big)
+\mathcal O(\upsilon^3)\,.
\label{eq:AdS2HABC}
\end{eqnarray}
It remains to verify that this is indeed the correct form of the supergeometry, i.e. that the above expressions solve the supergravity constraints to the relevant order. This is not guaranteed since they were derived from the flatness of a Lax connection which we simply postulated. It is known that it is enough to demonstrate that the Green-Schwarz string action possesses kappa-symmetry in order to say that the background is a solution to the supergravity equations. This in turn simply amounts to verifying that the components of the torsion $T^A$ and NS--NS superfield strength $H$ have the appropriate components involving fermionic supervielbeins whose form is dictated by the supergravity constraints \eqref{TA} and \eqref{eq:H}.
Indeed, with the above choice of the spin connection \eqref{eq:AdS2Omega} it is not hard to verify that the type IIA supergravity torsion constraint \eqref{TA} is satisfied to the quadratic order in $\upsilon$. As far as the constraint \eqref{eq:H} is concerned we should substitute into it the expressions \eqref{eq:AdS2vielbeins} and \eqref{eq:AdS2HABC} and check that the resulting superform $H$ is closed. This can indeed be verified but the calculation is somewhat lengthy and we leave the details for Appendix \ref{sec:dH}.
Thus we have constructed the supergeometry of the $AdS_2\times S^2\times T^6$ solution of type IIA supergravity to all orders in the 8 supercoset fermions $\vartheta$ and to the second order in the 24 non--supercoset fermions $\upsilon$. Although in this paper we have not obtained the RR superfield strengths $F_2$ and $F_4$, which are needed for the construction of D--brane actions in this superbackground, these can be found using the Bianchi identities, the corresponding superspace constraints given in \cite{Grassi:2009yj} and the above form of the supervielbeins.
By constructing the zero--curvature Lax connection we have demonstrated that the GS string in $AdS_2\times S^2\times T^6$ is classically integrable up to quadratic order in the non--supercoset fermions and to all orders in the coset fermions thus extending the results of \cite{Sorokin:2011rr} which considered all the fermions to the second order only.
\section{Properties of the Lax connection}
\subsection{$\mathbb Z_4$--invariance}
The Lax connection \eqref{eq:Lax-connection} with $L_{coset}$ and $L'$ given, respectively, in \eqref{Lcoset} and \eqref{L'} is invariant under the $\mathbb Z_4$--transformations \eqref{Z4auto} of the isometry generators and the inversion of the spectral parameter \eqref{Ox} and \eqref{Oalpha}. This demonstrates that the contribution of the non--coset fermions does not spoil the $\mathbb Z_4$--symmetry of the \emph{supercoset} Lax connection which is of crucial importance for the derivation of the algebraic curve and the Bethe ansatz equations, both classical and quantum \cite{Beisert:2005bm,Gromov:2008bz,Babichenko:2009dk,SchaferNameki:2010jy}. The $\mathbb Z_4$--invariance induces the corresponding conjugation symmetry of the monodromy matrix of the Lax connection
$$\Omega^{-1}\, M(1/{\tt x})\, \Omega = M({\tt x})$$
used for the construction of the algebraic curve \footnote{We thank Kostya Zarembo for the discussion of these points.}.
\subsection{The Lax connection and conserved currents}
Let us now present an interesting observation how the Lax connection $L=L_{coset}+\alpha_2L'$ can be related to the conserved current associated with the superisometries. First of all notice that in the limit $\alpha_2=\epsilon\rightarrow0$, in which
\begin{equation}\label{eps}
\alpha_1=\frac{1}{2}\epsilon^2+\mathcal O(\epsilon^4)\,,\qquad\beta_1=-\frac{1}{2}\epsilon+\mathcal O(\epsilon^2)\,, \qquad \qquad \beta_2=1+\mathcal O(\epsilon^2)\,,
\end{equation}
and
$$
V=\beta_2+\beta_1\Gamma_{11}\rightarrow 1, \qquad V^\dagger=\beta_2-\beta_1\Gamma_{11}\rightarrow 1\,,
$$
the Lax connection reduces to
\begin{equation}\label{Leps}
L=K+\mathcal O(\epsilon)\,,
\end{equation}
where $K(X,\vartheta)$ is the supercoset Cartan form introduced in \eqref{K}.
In fact, the term denoted by $\mathcal O(\epsilon)$ in the (gauge--transformed) Lax connection is the worldsheet Hodge dual of a superstring conserved current $J$ associated with the background superisometries, namely
\begin{equation}\label{J}
*J=\lim_{\epsilon\rightarrow0}\,\frac{1}{\epsilon}\,(gLg^{-1}-dgg^{-1})= g\lim_{\epsilon\rightarrow0}\frac{L-K}{\epsilon}g^{-1}\,,
\end{equation}
where $g(X,\vartheta)$ is the superisometry group element determining $K$ in \eqref{K}.
The conservation of $J$, i.e. $d*J=0$, follows from the flatness of the Lax connection and the Cartan form
\begin{equation}
dL-LL=0\,,\qquad dK-KK=0\,.
\end{equation}
Indeed, in view of \eqref{Leps}, we have
\begin{eqnarray}
d*J&=&
g\lim_{\epsilon\rightarrow0}\frac{dL-dK}{\epsilon}g^{-1}
-*JgKg^{-1}
-gKg^{-1}*J
\nonumber\\
&=&
g\lim_{\epsilon\rightarrow0}\frac{(L-K)K+K(L-K)}{\epsilon}g^{-1}
-*JgKg^{-1}
-gKg^{-1}*J
=0\,.
\end{eqnarray}
Note that in the case of the supercoset Lax connection \eqref{Lcoset} (when $\upsilon=0$), the current constructed in this way coincides with the conserved current of the $G/H$ sigma--model considered in \cite{Bena:2003wd,Arutyunov:2008if,Sorokin:2010wn}
\begin{equation}\label{Jcoset}
J_{coset}=g(E^AP_A-\frac{1}{2}Q\Gamma_{11}*E)g^{-1}\,.
\end{equation}
We can now write the correction \eqref{L'} to the Lax connection in terms of (transformed) components of the conserved current as
\begin{equation}
L'=g^{-1}*(\tilde J-\tilde J_{coset})g\,,
\end{equation}
where $\tilde J$ and $\tilde J_{coset}$ are, respectively, the complete conserved current (to second order in $\upsilon$) \eqref{J} and the conserved current of the supercoset model \eqref{Jcoset}, and the tilde means that in their expressions we perform the following substitutions of spinorial quantities
$$E\rightarrow V^\dagger E,\qquad \nabla\upsilon\rightarrow V^\dagger\nabla\upsilon \qquad {\rm and} \qquad\upsilon\rightarrow V\upsilon.$$
For instance,
\begin{equation}
\tilde J_{coset}=g(E^AP_A-\frac{1}{2}Q\Gamma_{11}V^\dagger*E)g^{-1}\,.
\end{equation}
Whether this fact is of some deeper significance remains to be understood. Perhaps, a better understanding of this could lead to a proposal for the complete Lax connection to all orders in the non--coset fermions. We leave this problem for future analysis.
\subsection{Relation to the Lax connections constructed in \cite{Sorokin:2010wn,Sorokin:2011rr}}
In \cite{Sorokin:2010wn,Sorokin:2011rr} Lax connections for the superstring in $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ have been constructed (up to second order in fermions) using components of the Noether currents of the corresponding superisometries $OSp(6|4)$ and $PSU(1,1|2)\times U(1)^6$ (note that only the Abelian $U(1)^6$--currents of the $E(6)$ isometries enter the Lax connection)
\begin{equation}\label{Jcurrent}
J=J_{\mathcal B}+J_{susy}\,,
\end{equation}
where the conserved current $J_{\mathcal B}$ of the bosonic isometries has the form
\begin{equation}\label{JB}
J_{\mathcal B}=J_1+J_2\,,
\end{equation}
\begin{eqnarray}\label{J12}
J_1&=&\Big(e^A(X)+i\Theta\Gamma^A {\nabla}\Theta+i\Theta\Gamma^A\Gamma_{11}*\nabla\Theta\nonumber\\
&&-\frac{2}{R}e^B\,\Theta\Gamma^A\mathcal P\gamma_\star \Gamma_B\Theta
-\frac{2}{R}*e^B\,\Theta\Gamma^A\mathcal P\gamma_\star\Gamma_{11} \Gamma_B\Theta\,\Big)(gP_A\,g^{-1})|_{\vartheta=0},\nonumber\\
\\
J_2&=&\frac{i}{8}\left(e^C\,\Theta\Gamma^{AB}{}_C\Theta
-\ast e^C\,\Theta\Gamma^{AB}{}_C\Gamma_{11}\Theta\right)R_{AB}{}^{DE}\,(g\,M_{DE}\,g^{-1})|_{\vartheta=0}\hspace{50pt}\nonumber
\end{eqnarray}
and the supersymmetry current $J_{susy}$ is
\begin{equation}\label{Jsusy}
J_{susy}=\frac{i}{R}\,(g Q g^{-1})|_{\vartheta=0}\,\gamma_\star\,(\ast
e^A\,\Gamma_A\Gamma_{11}\Theta-e^A\,\Gamma_A\Theta )\,.
\end{equation}
In \eqref{J12} and \eqref{Jsusy} the isometry group element \eqref{K} is evaluated at $\vartheta=0$ \footnote{Note that $K_A(X)=(gP_A\,g^{-1})|_{\vartheta=0}$ and $\Xi(X)=\gamma_\star(gQ g^{-1})|_{\vartheta=0}$ are simply the Killing vector and the Killing spinor of the superisometries (see \cite{Sorokin:2010wn,Sorokin:2011rr} for more details).}.
The Lax connection constructed in \cite{Sorokin:2010wn,Sorokin:2011rr} has the following form
\begin{equation}\label{Lambda}
\Lambda=\alpha_1\,J_1|_{\Theta=0}+\alpha_2*J_1+\alpha_2^2\,J_2+\alpha_2(1+\alpha_1)*J_2-\alpha_2(\beta_1 J_{susy}-\beta_2*J_{susy})\,,
\end{equation}
where the coefficients $\alpha_{1,2}$ and $\beta_{1,2}$ are the same as in \eqref{Lcoset}.
Now note that in the limit \eqref{eps} the Lax connection reduces to the Hodge dual of the conserved current \eqref{Jcurrent}
\begin{equation}\label{e0}
\lim_{\epsilon\rightarrow 0}\, \frac{1}{\epsilon}\Lambda=*J\,.
\end{equation}
Comparison of eq. \eqref{e0} with \eqref{J} suggests a non--straightforward relation between the two connections via the following gauge transformation depending on the spectral parameter and accompanied by the shift in the $X$--dependence of $\Lambda$ \footnote{The shift $X\rightarrow X+i\upsilon\Gamma^A\vartheta$ in $\Lambda$ is required since $\Lambda$ and $L$ are constructed in different coordinate systems (see Section 2 of \cite{Cagnazzo:2009zh} for more details about the choice of the coordinate basis).}
\begin{equation}\label{LLambda}
\Lambda(X^M+i\upsilon\Gamma^M\vartheta,\Theta)=\mathcal G_{\mathbf x}(X,\Theta)\,L(X,\Theta)\,\mathcal G_{\mathbf x}^{-1}(X,\Theta)-d{\mathcal {G_{\mathbf x}G_{\mathbf x}}}^{-1}(X,\Theta)\,,
\end{equation}
where both sides are truncated to quadratic order in fermions, $\Gamma^M=\Gamma^A\,e_A{}^M(X)$ and $\mathcal G_{\mathbf x}(X,\Theta)$ is an isometry supergroup element depending on the spectral parameter $\mathbf x$, which in the exponential parametrization has the following form
\begin{equation}\label{gvtheta}
\mathcal G_{\mathbf x}(X^A,\Theta)=\,e^{(X^A+i\upsilon\Gamma^A(1-V^2)\vartheta) P_A}\,e^{Q V\vartheta\,}h(i_\Delta\,\Omega_0^{AB})\,,
\end{equation}
where
$$
h(i_\Delta\,\Omega_0^{AB})=e^{-\frac{1}{2}i\upsilon\Gamma^C(1-V^2)\vartheta\,\Omega_{0C}{}^{AB}(X,\vartheta) M_{AB}}
$$
is a compensating gauge transformation in the stability subgroup $H$ (\emph{i.e.} $SO(1,3)\times U(3)$ or $SO(1,1)\times SO(2)$) of the superisometry group $G$.
\if{}
\\
\\
\textbf{SKETCH OF THE PROOF OF \eqref{LLambda}. }First we should check that \eqref{LLambda} holds for terms with $\nabla\Theta$. To this end it is enough to consider the gauge transformation of
$$
L=(1+\alpha_1)e^A(X)P_A+\alpha_2* e^A(X)P_A+(2i\alpha_2\upsilon\Gamma^A*\nabla\vartheta+2i\alpha_2\upsilon\Gamma^A\Gamma_{11}\nabla\vartheta)P_A+\cdots\,.
$$
Applying the $e^{(X+\Delta)P}$ part of \eqref{gvtheta} to the above terms we get
\begin{equation}\label{1}
\hat L=(1+\alpha_1)e^A(X)K_A(X+\Delta)-e^A(X+\Delta)K_A(X+\Delta)+\alpha_2* e^A(X)K_A(X+\Delta)+(2i\alpha_2\upsilon\Gamma^A*\nabla\vartheta+2i\alpha_2\upsilon\Gamma^A\Gamma_{11}\nabla\vartheta)K_A+\cdots\,
\end{equation}
where $\Delta=i\upsilon\Gamma^A(1-V^2)\vartheta e_A^M(X)$. At this stage of the proof we do not need to consider the variation of $K_A(X+\Delta)$ and can just replace it with $K_A(X)$. Then eq. \eqref{1} takes the form
\begin{eqnarray}\label{2}
&\hat L=\alpha_1e^A(X)K_A-\nabla(\upsilon\Gamma^A(-\alpha_1+\alpha_2\Gamma_{11})\vartheta)K_A+\alpha_2* e^A(X)K_A+(2i\alpha_2\upsilon\Gamma^A*\nabla\vartheta+2i\alpha_2\upsilon\Gamma^A\Gamma_{11}\nabla\vartheta)K_A+\cdots\,&\nonumber\\
&=\alpha_1(e^A(X)+i\nabla(\upsilon\Gamma^A\vartheta))K_A+\alpha_2(* e^A(X)+2i\upsilon\Gamma^A*\nabla\vartheta)K_A+i\alpha_2(\Theta\Gamma^A\Gamma_{11}\nabla\Theta)K_A+\cdots\,&\\
&=\alpha_1e^A(X^M+i\upsilon\Gamma^M\vartheta)K_A+\alpha_2(* e^A(X^M+i\upsilon\Gamma^M\vartheta)+i\Theta\Gamma^A*\nabla\Theta +i\Theta\Gamma^A\Gamma_{11}\nabla\Theta)K_A+\cdots\,&\nonumber\\
&=\Lambda(X^M+i\upsilon\Gamma^M\vartheta,\Theta)&\nonumber
\end{eqnarray}
\fi
Actually, as an independent derivation procedure, alternative to that described in Section \ref{comleteL}, we also got the form of the terms in $L'$ quadratic in fermions in the Lax connection \eqref{L'} by performing the inverse gauge transformation, from $\Lambda$ to $L$.
Note that in contrast to \eqref{Lcoset} and \eqref{L'} the Lax connection \eqref{Lambda} is not directly invariant under the $\mathbb Z_4$--transformations \eqref{Z4auto}, \eqref{Ox} and \eqref{Oalpha}. In particular, its first ($\alpha_1$--dependent) term acquires the shift $-2e^{A}\Omega(gP_Ag^{-1}|_{\vartheta=0})$. To get back $\Lambda$ in its initial form the $\mathbb Z_4$--transformed Lax connection
$$
\Omega(\Lambda({\tt x}))=\Omega^{-1}\Lambda(\frac{1}{\tt x})\, \Omega
$$
should undergo a compensating gauge transformation $\mathcal G_\Omega$ and one finds
$$
\Lambda=\mathcal G_\Omega\,\Omega(\Lambda)\,\mathcal G^{-1}_\Omega-\mathcal G_\Omega d\mathcal G^{-1}_\Omega, \qquad {\rm where} \qquad \mathcal G_\Omega=\mathcal G_{\mathbf x}\Omega^{-1} \mathcal G_{\mathbf x}^{-1}\Omega\,,
$$
$\mathcal G_{\mathbf x}(X,\Theta)$ is the same as in \eqref{LLambda} and $\Lambda$ is evaluated at $X^M+i\upsilon\Gamma^M\vartheta$. Of course, this gauge transformation, which also affects the spectral parameter $\tt x$, is nothing but a different form of the relation \eqref{LLambda} taking into account the $\mathbb Z_4$--invariance of $L$.
\subsection{Lax connection and kappa--symmetry}
The Green--Schwarz formulation of the superstring is invariant under the local fermionic transformations of the target--space coordinates $Z^{\mathcal M}=(X^M,\Theta^\mu)$ which satisfy the following properties
\begin{equation}\label{kappastring}
\delta_\kappa Z^{\mathcal M}\,{\mathcal E}_{\mathcal M}{}^{ \alpha}=
\frac{1}{2}(1+\Gamma)^{ \alpha}_{~ \beta}\,
\kappa^{ \beta}(\xi),\qquad { \alpha}=1,\cdots, 32
\end{equation}
\begin{equation}\label{kA}
\hskip-2.5cm\delta_\kappa Z^{\mathcal M}\,{\mathcal E}_{\mathcal M}{}^A=0,
\qquad A=0,1,\cdots,9
\end{equation}
where $\kappa^{ \alpha}(\xi)$ is a 32--component spinor
parameter, $\frac{1}{2}(1+\Gamma)^{
\alpha}_{~ \beta}$ is a spinor projection matrix with
\begin{equation}\label{gbs}
\Gamma=\frac{1}{2\,\sqrt{-\det{g_{ij}}}}\,\epsilon^{ij}\,{\mathcal
E}_{i}{}^A\,{\mathcal E}_{j}{}^B\,\Gamma_{AB}\,\Gamma_{11}, \qquad
\Gamma^2=1\,,
\end{equation}
and $g_{ij}$ is an induced worldsheet metric.
The string equations of motion \eqref{Psi} and \eqref{B} transform into each other under the kappa--symmetry variations. Since the condition for the Lax connection to have zero--curvature is in one to one correspondence with the equations of motion, it is natural to assume that on the mass--shell the Lax connection should be invariant under the kappa--symmetry transformations, at least, modulo a gauge transformation. This is indeed so in the case of the supercoset sigma--models (see e.g. \cite{Arutyunov:2008if}). The explicit check that also the non--coset Lax connection \eqref{eq:Lax-connection}, \eqref{L'} possesses this property would be somewhat cumbersome, but fortunately one should not do this, because there is a simple generic proof that makes this fact evident. Indeed, since any Lax curvature depends on the left--hand--sides of the equations of motion (as \emph{e.g.} in \eqref{R01} and \eqref{eq:dLAdS2}), its variation under \eqref{kappastring} and \eqref{kA} also depends on the field equations and hence vanishes on--shell. This means that kappa--variation of the Lax connection leaves its curvature zero and, therefore, the kappa--transformed Lax connection is related to the initial one by a corresponding infinitesimal gauge transformation taking values in the isometry superalgebra.
\section{Conclusion}
We have constructed the zero--curvature Lax connections for Green--Schwarz superstrings in $AdS_4\times CP^3$ and $AdS_2\times S^2\times T^6$ superbackgrounds which generalize the corresponding supercoset sigma--model Lax connections with contributions due to the physical world--sheet fermionic modes associated with non--supersymmetric directions of the target superspaces. We have shown that the contribution of the non--coset fermions does not spoil the important property of the Lax connections being $\mathbb Z_4$--invariant and demonstrated how the obtained Lax connections are related via gauge transformations to the Lax connections constructed in \cite{Sorokin:2010wn,Sorokin:2011rr} with the use of an alternative (Noether--current) prescription.
Having at hand Lax connections which include the contribution of non--coset worldsheet modes one can address the problem of how these modify the algebraic curve and Bethe ansatz equations for the full superstring theory in these backgrounds. This should lead to a more general approach to integrability of Green--Schwarz superstrings which does not rely on having a supercoset sigma--model description of the string.
The terms in the Lax connections containing the non--supercoset fermions $\upsilon$ have been computed to the second order in $\upsilon$. An interesting and important open problem is to understand the structure of the Lax connections to all the orders in the non--coset fermions. Presumably, the series in $\upsilon$ (eq. \eqref{L'}) would converge into covariant expressions in terms of background superfields (supervielbeins, connection etc.) as happens for the supercoset Lax connections expressed in terms of the superisometry Cartan forms.
In the process of the construction of the Lax connections we have obtained the form of the superfield quantities (supervielbeins, connection, NS--NS field strength and dilatino) that describe the $AdS_2\times S^2\times T^6$ superbackground to all orders in the supercoset fermions and to the second order in the non--coset ones. We have also obtained the explicit form of the spin connection of the $AdS_4\times CP^3$ superspace (to the second order in $\upsilon$) which was left out in \cite{Gomis:2008jt}. In contrast to the Green--Schwarz formulation, the knowledge of the form of the spin connection is required, for instance, for the pure spinor description of superstrings in curved superbackgrounds \cite{Berkovits:2001ue} and, in particular, is needed for extending to the full $AdS_4\times CP^3$ superspace the supercoset pure--spinor sigma--model of \cite{Fre:2008qc,Bonelli:2008us,D'Auria:2008cw}. With some more efforts, which will be made elsewhere, one can also compute the form of the superfield strengths $F_2$ and $F_4$ of the RR fluxes in type IIA $AdS_2\times S^2\times T^6$ superbackgrounds and corresponding quantities describing this superbackground in the type IIB case. These are also required for the construction of the pure spinor string action and for studying D--branes in these superbackgrounds.
Finally a detailed knowledge of string theory in $AdS_2\times S^2\times T^6$ might shed light on the corresponding $AdS_2/CFT_1$ holographic duality which so far has not been well understood. This correspondence is especially important due to the relation to black holes in $D=4$ that have an $AdS_2\times S^2$ near--horizon geometry. The integrable string model considered here could for example be used to make predictions for anomalous dimensions of operators on the gauge--theory side which could be compared to those computed in a given candidate dual theory.
\subsection*{Acknowledgements}
The authors are grateful to Igor Bandos, Arkady Tseytlin and Konstantin Zarembo for valuable discussions and comments. Work of A.C. and D.S. was partially supported by the INFN Special Initiative TV12. D.S. was also supported in part by the MIUR-PRIN contract 2009-KHZKRX and the grant FIS2008-1980 of the Spanish MICINN. D.S. is grateful to the Department of Theoretical Physics of the Basque Country University for hospitality and the IKERBASQUE Foundation for a visiting fellowship. The research of L.W. was supported in part by NSF grants PHY-0555575 and PHY-0906222.
\newpage
|
\section{Introduction}
The Dirac operator represents the effects of the quark fields on the gauge field.
The determinant of this operator appears in the QCD partition function. This operator
depends on the gauge field, we are looking for the statistics of its eigenvalues.
The matrix of the Dirac operator is very large, its linear dimension scales with the
lattice volume. The inverse of the Dirac operator occurs in physically measurable
quantities. Therefore the low eigenvalues of the Dirac operator are very important.
\\
In the chirally broken phase (below $T_{c}$) the statistics of the Dirac spectrum
can be described by Random Matrix Theory \cite{Verbaarschot:2000dy}. However, we want to
analyze the statistics of the Dirac eigenvalues above $T_{c}$ in the chirally symmetric phase.
In this case we do not know any model for describing the statistics of the Dirac eigenvalues.
In the first step we regard the Dirac operator as a general fluctuating matrix with the same
global symmetry as QCD. From this viewpoint we have two extreme
possibilities:
\begin{itemize}
\item Random Matrix type spectrum
\\
In this case typical fluctuations in the matrix elements can freely mix eigenvectors,
the eigenvectors are extended. If we increase the volume of the lattice, then the spatial extension
of these eigenmodes will increase too which means that they are delocalized. They fill almost the entire
lattice volume. The eigenvalues have Random Matrix statistics, they are not statistically independent.
\item Poisson type spectrum
\\
In this case fluctuations in the matrix elements cannot mix eigenvectors, the eigenvectors
are localized. The spatial extension of these eigenmodes will not depend on the volume of the lattice.
These eigenmodes are localized to a certain place in the lattice.
The eigenvalues follow Poisson statistics in this case, they are statistically independent.
\end{itemize}
Now the question is whether the eigenmodes of the Dirac operator are delocalized Random Matrix type,
or localized Poisson type.
\\
We first summarize the recent results that appeared in the literature.
Above $T_{c}$ in the chirally symmetric phase the spectral density of the Dirac operator around zero
vanishes \cite{Banks:1979yr} and Random Matrix Theory has predictions at such soft edge \cite{Forrester}.
Lattice simulations did not find agreement with these predictions. On the other hand, they indicate
bulk Random Matrix statistics for the full Dirac spectrum \cite{Pullirsch:1998ke}. This would implicate
that the eigenmodes of the Dirac operator are delocalized. However, in Ref.\ \cite{GarciaGarcia:2006gr}
the authors showed that around $T_{c}$ the eigenvalue statistics changes from Random Matrix type
to Poisson type spectrum. In Ref.\ \cite{Kovacs:2009zj} the author argued that the first few eigenvalues
of the Dirac operator are localized.
\\
\section{Simulation details}
We examine the Dirac spectrum by computing the first few hundred eigenvalues of the Dirac operator for
each gauge configuration. We divided this spectral range into many parts, and we examined the eigenvalue
statistics separately in each part of the spectrum. We performed this analysis in SU(2) quenched theory in
\cite{Kovacs:2010wx}. In that case we saw a transition between localized and delocalized eigenmodes in the
spectrum. We want to know if this transition also appears in QCD. For this reason we must include
dynamical fermions and we have to simulate SU(3) gauge theory. We examine whether the determinant of the
Dirac operator modifies the statistics of the low Dirac eigenvalues or not. Our data are based on SU(3)
gauge theory with (2+1) flavors of dynamical quarks at physical quark masses. We set the temperature
to about $2.6T_{c}$, well in the chirally symmetric phase. We use the action of Budapest-Wuppertal group
\cite{Aoki:2006br}. This involves the staggered Dirac operator with two levels of stout smearing and a
Symanzik improved gauge action.
We have lattices with two different lattice spacings and for each lattice spacing we have gauge configurations on several
volumes. For the details see Table~\ref{tab:extension}.
\begin{table}
\caption{\label{tab:extension}The extension in time, and spatial direction of our lattices}
\begin{center}
\begin{tabular}{c | c c c}
$N_{t}$
&
\multicolumn{3}{c}{$N_{s}$} \\
\hline
$4$
&16 &24 & 32\\
$6$
& 24 & 36
\\
\end{tabular}
\end{center}
\end{table}
\section{Results}
Our main question is, whether the eigenmodes of the Dirac operator are localized, or delocalized.
To answer this question we have to know how to measure the spatial extension of an eigenmode.
We measure the volume of a general eigenmode with the help of this formula
\begin{equation}
{\cal V}=\left[\sum_{x:~lattice~site}\vert\psi_{i}\left(x\right)\vert^{4}\right]^{-1}
\label{eq:em_vol}
\end{equation}
This is equivalent to the product of the participation ratio and the total volume of the lattice.
To illustrate the usefulness of eq.\ \ref{eq:em_vol} we give a simple example. Let us have a normalized
eigenmode $\psi_{i}\left(x\right)$ which spreads out uniformly in a volume ${\cal V}$. At a given site
$x$ the absolute value squared of the wave function is equal to $\frac{1}{{\cal V}}$.
A short calculation shows that in this special case the quantity defined in eq.\ \ref{eq:em_vol} yields
the volume occupied by $\psi_{i}(x)$.
Now the question is how this volume scales with the linear extension of the
lattice. Our first guess is that this is proportional to the $d$-th power of $L$, the lattice linear
extension
\begin{equation}
{\cal V}\left(L\right)=C \cdot L^{d},
\end{equation}
where $d$ is the effective dimension of the eigenmode. Now we will examine this quantity throughout
the spectrum. In Fig.\ \ref{fig:em_dim} we display this effective dimension as a function of the
eigenvalue. The eigenvalues are in lattice units and $d$ is the effective dimension of the average
eigenvector corresponding to a certain eigenvalue range. At the low end of the spectrum this quantity
is equal to zero which means that these eigenmodes are localized, their spatial extension does not
change as we increase the lattice volume. Going up in the spectrum the effective dimension increases.
As we reach the bulk of the spectrum the dimension becomes roughly $3$ which means
that the bulk eigenvectors spread out in all three spatial directions, they are completely delocalized.
Between these two extremes there must be some transition.
\begin{figure}
\begin{center}
\includegraphics[width=0.6\columnwidth,keepaspectratio]{Figs/Effective_dimension.pdf}
\caption{\label{fig:em_dim}Effective dimension of the eigenvectors, $N_{t}=4$}
\end{center}
\end{figure}
\\
We examine the other quantity that determines how freely the eigenmodes can mix, the spectral density.
If the number of eigenmodes in a given eigenvalue range is small and these eigenmodes are localized,
they hardly mix. On the other hand if the number of eigenmodes is large in the same range, and the
eigenmodes are delocalized, then they can freely mix. We want to measure both of these effects. Therefore,
we define a quantity that we call the cumulative volume fill fraction (CVFF). This is the sum of the volumes of all the eigenmodes up to a given eigenvalue
$\lambda$, normalized by the total number of configurations. We measure this quantity in box size units.
We display the CVFF in Fig.\ \ref{fig:cum_fill}.
At the low end of the spectrum this quantity is much smaller than unity. Therefore we expect, that these
eigenmodes are independent, and they are produced independently in different subvolumes. Near
$\lambda a=0.4$ this quantity becomes much bigger than unity. This means that these modes are strongly
overlap.
\begin{figure}
\begin{center}
\includegraphics[width=0.7\columnwidth,keepaspectratio]{Figs/cum324.pdf}
\caption{\label{fig:cum_fill}Cumulative volume fill fraction $N_{t}=4,N_{s}=32$}
\end{center}
\end{figure}
\\
\begin{figure}
\begin{center}
\begin{tabular}{cc}
\includegraphics[width=0.4\columnwidth,keepaspectratio]{Figs/0_15_0_19_324.pdf}&
\includegraphics[width=0.4\columnwidth,keepaspectratio]{Figs/0_29_0_32_324.pdf}\\
\includegraphics[width=0.4\columnwidth,keepaspectratio]{Figs/0_34_0_35_324.pdf}&
\includegraphics[width=0.4\columnwidth,keepaspectratio]{Figs/0_375_0_385_324.pdf}
\\
\end{tabular}
\end{center}
\caption{\label{fig:unfold}Unfolded level spacing distribution $N_{t}=4,N_{s}=32$,
$\Delta\lambda = \frac{\lambda_{n+1}-\lambda_{n}}
{\langle \lambda_{n+1}-\lambda_{n} \rangle}$}
\end{figure}
Next we examine the eigenvalue statistics separately in the four regions indicated in
Fig.\ \ref{fig:cum_fill}. In order to compare the eigenvalue statistics with Poisson and Random Matrix
type spectra we have to unfold the spectrum. After this transformation we retain only the universal
correlations in the spectrum. To this end we rescale the eigenvalues to make the spectral density
equal to unity throughout the whole spectrum. We have to choose some statistics which we can easily obtain from our data
and for which we have analytical prediction in the Poisson and Random Matrix case. We computed the
unfolded level spacing distribution (ULSD) of eigenvalues. For Poisson type localized eigenmodes the ULSD
follows a simple exponential. At the low end of the spectrum our result is in Fig.\ \ref{fig:unfold}(a).
We can see that our data match this exponential well, demonstarting that the eigenvalues are statistically
independent.
Going up in the spectrum the statistics changes. (Fig.\ \ref{fig:unfold}(b),Fig.\ \ref{fig:unfold}(c)).
For Random Matrix type delocalized eigenmodes the ULSD can be described by the Wigner surmise of the
corresponding random matrix ensemble. In the bulk of the spectrum in Fig.\ \ref{fig:unfold}(d) we can
see that our data match perfectly this Wigner surmise. In this case the eigenvalues are not statistically
independent, instead they repel each other. We see a transition from Poisson type localized eigenmodes to
Random Matrix type delocalized eigenmodes in this case too.
\\
\begin{figure}
\begin{center}
\includegraphics[width=0.4\columnwidth,keepaspectratio]{Figs/stag_deg_cont.pdf}
\caption{\label{fig:stag_deg_a}ULSD between the first two eigenvalues }
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.4\columnwidth,keepaspectratio]{Figs/Formdoublet.pdf}
\caption{\label{fig:form_d}Average level spacing of the first few eigenvalues}
\end{center}
\end{figure}
Zooming on the lowest part of the spectrum reveals that the statistics of the lowest
Dirac eigenvalues differs from the Poisson statistics. We examined the ULSD between the first two
eigenvalues (Fig.\ \ref{fig:stag_deg_a}). We saw that the first two eigenvalues
are not independent, instead they attract each other. Going to our finer lattices, we can see
that this effect becomes stronger. We do not see such an effect on our simulation with SU(2) quenched
gauge theory with the overlap Dirac operator. We want to know what is the cause of this effect. We used
staggered fermions, and we know that the eigenvalues of the staggered Dirac operator are four fold
degenerate in the continuum limit. We used two levels of stout smearing which reduces the mixing between
high frequency modes of the theory. Therefore the spectrum of the Dirac operator looks more
"continuum-like". In this case the staggered eigenvalues start to form a doublets. This can easily be
seen in Fig.\ \ref{fig:form_d} where we plot the average level spacing corresponding to $n$-th eigenvalue. We see that there are two relevant scales in the problem:
\begin{itemize}
\item
splitting inside doublets
\item
spacing between doublets
\end{itemize}
\begin{figure}
\begin{center}
\includegraphics[width=0.4\columnwidth,keepaspectratio]{Figs/stag_deg_artefact.pdf}
\caption{\label{fig:last_plot}ULSD in the same range for different volumes}
\end{center}
\end{figure}
The splitting inside doublets does not depend on the volume, but the spacing between doublets
will decrease when we increase the volume. Therefore if the volume is large enough we cannot see
this effect at a certain lattice spacing, because if the spacing between doublets becomes
comparable to the splitting, we cannot see doublets any more. This effect can be seen
in Fig.\ \ref{fig:last_plot}.
Examining the level spacing distribution at the very low end of the spectrum for three different
volumes shows that this effect becomes weaker in larger volumes.
\section{Conclusion}
We have seen that the fermionic determinant does not destroy the transition in the spectrum from
Poisson type localized eigenmodes to Random Matrix type delocalized eigenmodes. We saw the transition
on our finer lattice too. We see that the transition remains also in the thermodynamic limit.
Because we use the staggered Dirac operator, the eigenvalue statistics on the low end of the spectrum is
disturbed by the doublets, but this effect is only a finite volume artefact.
\\
|
\section{Introduction}
It is generally taken on faith that the geometry underlying general relativity is bosonic. In quantum gravity, this faith is buttressed by the observation that low energy, small-amplitude excitations of the gravitational field are spin-2 gravitons, and should therefore be quantized as bosons. On the other hand, it is well known that certain constrained or otherwise non-linear field theories whose low energy excitations are bosonic, can nevertheless give rise to emergent structures in the non-perturbative regime with fermionic (or anyonic) statistics \cite{Arnsdorf:1998vq,Finkelstein:1968hy,Giulini:1993gd,Skyrme:1961vq,Skyrme:1962vh,Skyrme:Original,Skyrme:Original2,Solitons,TopologicalSolitons,Williams:Skyrme}. It is conceivable then that the phase space structure of general relativity could admit fermionic modes upon quantization.
To support this hypothesis, in a recent article we constructed a purely geometric theory with precisely this peculiar property \cite{Randono:2010cd,Randono:2011bb}. The model presented there is not general relativity, however it is a dynamical geometry theory.
Specifically, we considered a propagating torsion theory in flat Minkowski space, where the torsion is constrained in such a way that the local degrees of freedom of the model are described by a non-linear sigma model with target space $Spin(3,1)$. By formally mapping the model onto the Skyrme model for strongly interacting baryons, we showed that the system could be quantized in such a way that isolated torsional charges of even charge behave as bosons under rotations and exchanges, whereas odd charges behave as fermions.
The existence of these fermionic geometries leads one to question whether this could be a generic feature of dynamical geometries or simply a peculiarity of the particular model. One may raise the speculative objection, for example, that the possibility of more exotic non-pertubative statistics is novel to torsional theories and cannot necessarily be extrapolated to non-torsional geometries. In this paper I will give strong evidence to quell this objection.
Specifically I will show that a much more familiar geometry, namely de Sitter space itself and a class of generalized de Sitter-like geometries can be quantized fermionically. The geometric arena I will work in is the gauge formulation of gravity where geometry is described by a reductive Cartan connection on a $Spin(4,1)$-bundle \cite{Randono:Condensate,Randono:dSSpaces,Utiyama:1973nq,Utiyama:1980bp,West:1978Lagrangian,Wise:2009fu,Wise:MMGravity,Randono:Review}. Surprisingly, the underlying mathematics allowing for fermionic quantization is the same as for the Skyrme model.
Bosonic geometry has long been an implicit tenet of quantum gravity theories. The possibility of fermionic geometries will likely require significant rethinking of foundational issues in quantum geometry. Furthermore, de Sitter space plays an important role in cosmology as the expected ground state of a universe with a positive cosmological constant and potentially as the future asymptote of our universe. This promotes the question {\it ``Is geometry bosonic or fermionic?"} beyond the realm of the rhetorical and places it on uncomfortably familiar ground.
\section{Geometry from Cartan's perspective}
The mathematical arena that I will work in is the reformulation of Einstein-Cartan theory as a symmetry broken gauge theory. I will briefly review this formalism below, but I refer the reader to my review paper \cite{Randono:Review} for a more extensive presentation.
In the gauge formulation of gravity, geometry is characterized by a Cartan connection $\mathcal{A}$ taking values in a reductive Cartan algebra $\mathfrak{g}=\mathfrak{h}\oplus\mathfrak{p}$. The reductive split of the algebra into a stabilizer subalgebra $\mathfrak{h}$ and its complement $\mathfrak{p}$ is facilitated through the introduction \cite{Stelle:1979aj,Stelle:1979va} of a new field $V$. This field acts as a symmetry breaking field analogous to the Higgs, ``breaking'' the $\mathcal{G}$ symmetry of the principle $\mathcal{G}$-bundle down to a subgroup $\mathcal{H}$ which stabilizes $V$ at each point. The field itself takes values in a space isomorphic to the coset space $\mathcal{G}/\mathcal{H}$. Thus, in total the geometry is characterized by a specification of the pair $\{\mathcal{A},V\}$, modulo $\mathcal{G}$-gauge transformations. The breaking of the symmetry allows for a separation of the spin connection from the frame field corresponding to the reductive split $\mathfrak{g}=\mathfrak{h}\oplus \mathfrak{p}$, so that $\mathcal{A}=\omega \oplus \frac{1}{\ell}e$ where $\ell$ is a parameter with dimension of length. Here, $\omega$ is the $\mathfrak{h}$-valued spin-connection, and $e$ is the $\mathfrak{p}$-valued tetrad. For the case of gravity in $(3+1)$-dimensions with a positive cosmological constant, the gauge group is $\mathcal{G}=Spin(4,1)$, the double cover of the de Sitter group, and the stabilizer subgroup is $\mathcal{H}=Spin(3,1)$. The parameter $\ell$ can be related to the cosmological constant $\Lambda$ by $\ell=\sqrt{3/\Lambda}$.
To see this more explicitly, consider the adjoint representation of $\mathcal{G}=Spin(4,1)$. Let hatted-upper case Roman indices $\{\h{I}, \h{J}, \h{K},\dots\}$ range from $0$ to $4$, and unhatted indices $\{I,J,K,\dots\}$ range from $0$ to $3$. The symmetry breaking field $V^{\h{I}}$ is a vector representation of $Spin(4,1)$ whose magnitude is constrained by $\eta_{\h{I}\h{J}}V^{\h{I}}V^{\h{J}}=1$ with $\eta_{\h{I}\h{J}}=diag(-1,1,1,1,1)$. The tetrad can then be identified with $e^{\h{I}}\equiv \ell D_A V^{\h{I}}$ and the spin connection with $\omega^{\h{I}\h{J}}=\mathcal{A}^{\h{I}\h{J}}-2D_\mathcal{A} V^{[\h{I}} \,V^{\h{J}]}$.
One can always choose a specific gauge locally where $V^{\h{I}}=(0,0,0,0,1)$. For the rest of the paper we will refer to this gauge as the Einstein-Cartan (EC) gauge. In the EC gauge, the tetrad simplifies to $e^I=\ell \mathcal{A}^{I4}$, and the spin connection to $\omega^{IJ}=\mathcal{A}^{IJ}$. In a Clifford algebra notation which we will employ here (see Appendix A) the symmetry breaking field is given by $V=V_{\h{I}}\gamma^{\h{I}}$ with $\gamma^{\h{I}}=(\gamma^I,\gamma_5)$. The action of a $Spin(4,1)$ gauge transformation generated by $g=g(x)$ is given by $\{\mathcal{A},V\}\rightarrow \{g\mathcal{A} g^{-1}-dg\,g^{-1},gVg^{-1}\}$.
Corresponding to the split\footnote{It is convenient in the Clifford algebra notation to define $e=\frac{1}{2}\gamma_I\, e^I$ which pulls the $\gamma_5$ out of the expression in the decomposition. See Appendix A for more details.} $\mathcal{A}=\omega+ \frac{1}{\ell}\gamma_5 e$, the curvature also splits into
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\mathcal{F}_\mathcal{A}=\underbrace{R_{\omega}-\frac{1}{\ell^2} e\w e}_{\mathfrak{h}}\ \ + \ \ \underbrace{\gamma_5 \frac{1}{\ell}T}_{\mathfrak{p}}\,.
\eeq
The constant curvature ($R_{\omega}=\frac{1}{\ell^2} e\w e$), zero torsion ($T=0$) condition is then given succinctly by $\mathcal{F}_{\mathcal{A}}=0$. Thus, in the gauge framework of gravity, de Sitter space is described by a flat connection. This allows one to easily embed the de Sitter solution into a larger space, which is at the same time small enough to easily characterize topologically. For this reason, I will focus on the configurations consisting of the pair $\{\mathcal{A}, V\}$ where $\mathcal{A}$ is restricted to be a flat connection. I will refine this phase space shortly.
\subsection{Flat Cartan geometries}
Restrict attention to manifolds with the topology of de Sitter space, namely $M\simeq \mathbb{R}\times \mathbb{S}^3$. Since $\pi_1(M)=0$, all flat connection are ``gauge related" to the zero connection. This implies that there exists a $g:\mathbb{R}\times \mathbb{S}^3 \rightarrow Spin(4,1)$ such that $\mathcal{A}=-dg g^{-1}$. The geometry is therefore characterized by the pair $\{\mathcal{A},V\}=\{-dg\,g^{-1},V\}$. However, one can always transform {\it both} fields to a gauge where $\mathcal{A}=0$ and $V'=g^{-1}Vg$ thereby pushing all the geometric information into the symmetry breaking field itself. In this gauge, which I will refer to as the trivial gauge, the geometry is completely characterized by the map $V:\mathbb{R}\times \mathbb{S}^3\rightarrow \mathcal{G}/\mathcal{H} \simeq \mathbb{R}\times \mathbb{S}^3$.
The goal is to categorize the space of flat Cartan connections topologically. To this end, consider first the set of maps that are deformable to a time independent map $V:\mathbb{S}^3 \rightarrow \mathcal{G}/\mathcal{H}\simeq \mathbb{R}\times \mathbb{S}^3$. The homotopy class of maps of this sort fall into discrete classes characterized by $\pi_3(\mathcal{G}/\mathcal{H})=\mathbb{Z}$. The integer labelling the sector in which the map resides is the number of times the $\mathbb{S}^3$ of the range winds around the $\mathbb{S}^3$ of the domain, or the winding number of the map.
As shown in \cite{Randono:dSSpaces, Randono:Review} the $\mathbb{Z}$-sectors of the phase space can be constructed as follows. Define $\ou{m}{h}{n}$ to be the time-independent map $\ou{m}{h}{n}:\mathbb{S}^3 \rightarrow Spin(4,1)$ given by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\ou{m}{h}{n}=\left[ \begin{matrix} h^m & 0 \\ 0 & h^n \end{matrix} \right] \quad \mbox{with} \quad h=X^4 \bm{1}+X^i \,i\sigma_i \label{Zgenerator}
\eeq
where $X$ denotes the usual embedding of the three sphere in $\mathbb{R}^4$ given explicitly in three-dimensional polar coordinates by
\beqa
X^1 &=& \sin\chi \,\sin{\theta} \,\cos\phi \nn\\
X^2 &=& \sin\chi \, \sin{\theta}\,\sin{\phi} \nn\\
X^3 &=& \sin\chi \, \cos{\theta} \nn\\
X^4 &=& \cos{\chi}
\end{eqnarray}}\newcommand{\w}{\wedge}\newcommand{\del}{\nabla}\newcommand{\ts}{\textstyle
The field $h$ is a map $h:\mathbb{S}^3 \rightarrow SU(2)$ with winding number one, and as such it is the generator of $\pi_3(SU(2))=\mathbb{Z}$. Since $Spin(4) \simeq SU(2)_\uparrow \times SU(2)_\downarrow$ and $Spin(4)$ is a subgroup of $Spin(4,1)$, $\ou{m}{h}{n}(x)\in Spin(4,1)$ and it has winding number $m+n$. To add back in the time dependence of the de Sitter solutions, first time-translate $\ou{m}{h}{n}$ to $\ou{m}{g}{n}=h_t \ou{m}{h}{n}$ where $h_t=\exp(\frac{1}{2}t \gamma_5 \gamma^0)$. A class of topologically distinct Cartan geometries is then given, in the EC gauge, by the configurations
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\{\mathcal{A},V\}=\{\ou{m}{\mathcal{A}}{n} \equiv-d\ou{m}{g}{n}\,\ou{m}{g}{n}{}^{-1}, \gamma_5\}\,. \label{AConfig}
\eeq
To understand the topological structure of these configurations more thoroughly it is useful to transform to the trivial gauge where $\mathcal{A}=0$. In this gauge, it can be shown that $V'=\ou{m}{V}{n}\equiv \ou{m}{g}{n}{}^{-1} \gamma_5 \ou{m}{g}{n}{}^{-1}$ viewed as a map $\ou{m}{V}{n}:\mathbb{R}\times \mathbb{S}^3 \rightarrow \mathcal{G}/\mathcal{H}$ has winding number $-q$ where $q\equiv m-n$. Thus, the integer $-q$ delineates the topological sectors of the space of flat Cartan connections corresponding to the homotopy group $\pi_3(\mathcal{G}/\mathcal{H})$.
Geometrically these configurations have a very simple interpretation. The metric induced from the pair $\{\ou{m}{\mathcal{A}}{n},V=\gamma_5\}$ is given by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\ou{m}{\bm{g}}{n}=-dt^2 +\cosh^2(t/\ell)\left(|q|^2 d\chi^2 +\sin^2(|q| \chi) \left(d\theta^2+\sin^2\theta \,d\phi^2 \right) \right)\,. \label{GConfig}
\eeq
For $|q| >0$, this metric describes a ``string of pearls" geometry consisting of $|q|$ copies of de Sitter space attached at their poles by regions where the metric becomes degenerate (see Fig. \ref{3Spheres}). The geometry of each individual 3-sphere is identical to that of de Sitter space itself aside from the poles. One can show that the volume of the three-sphere defined at the throat of de Sitter space at $t=0$ is given by $2\pi^2 \ell^3 (-q)$.
\begin{wrapfigure}{r}{0.5\textwidth}
\begin{center}
\includegraphics[width=0.4\textwidth]{3Spheres.jpg}
\end{center}
\caption{\label{3Spheres} A visualization of the generalized de Sitter space corresponding to $|q|=3$ on a constant time slice. On the left is the standard visualization where the radius is weighted by volume. The geometry consists of a string of $|q|$ three-spheres attached at the poles along two-spheres where the metric becomes degenerate. On the right is a more topologically accurate picture where the degenerate surfaces are pictured as extended regions denoted by dotted lines.}
\end{wrapfigure}
It should be stressed that all these configurations are solutions to the Einstein Cartan field equations, despite being degenerate \cite{Randono:dSSpaces, Randono:Review}. This serves to generalize de Sitter space, and allows one to embed the pure de Sitter solution into a larger phase space that is still manageable to work with.
\subsection{$\mathbb{Z}_2$ degeneracy of the phase space}
I will now show that the phase space containing the de Sitter configuration admits one further degeneracy. But first it is necessary to refine and define the phase space more precisely. Consider the set of flat Cartan geometries that asymptotically approach the pair $\{\ou{m}{\mathcal{A}}{n},V=\gamma_5\}$ in the past and future. In the trivial gauge this can be simplified to the set of maps $V:\mathbb{R}\times \mathbb{S}^3 \rightarrow \mathcal{G}/\mathcal{H}$ that asymptote to $\ou{m}{V}{n}$. In this fixed gauge, the only remaining gauge invariance is a global action of a constant $g\in Spin(4,1)$. Rather than modding out by diffeomorphism equivalence classes, we will consider the action of diffeomorphisms on the phase space that preserves the geometry on the boundaries. This allows for a classification of the action of diffeomorphisms into types categorized by topological properties. Thus, I will take the phase space $\mathcal{Q}$ to be the space of flat Cartan connections modulo the set of {\it identity connected}, boundary isometries.
The two lowest non-zero homotopy groups of the target space of $V$ are $\pi_3(\mathcal{G}/\mathcal{H})=\pi_3(\mathbb{S}^3)=\mathbb{Z}$ and $\pi_4(\mathcal{G}/\mathcal{H})=\pi_4(\mathbb{S}^3)=\mathbb{Z}_2$. I have already shown that the former is the topological property that allowed for the construction of a class of configurations labelled by the integer $-q$. The remaining $\mathbb{Z}_2$ degeneracy I will argue allows for fermionic quantization of de Sitter space. In total, the space of flat Cartan connections $\mathcal{Q}$ splits into topologically distinct sectors labelled by the winding number, and the $\mathbb{Z}_2$ degeneracy.
\section{The action of diffeomorphisms on $\mathcal{Q}$}
The goal now is to construct a generator of the $\mathbb{Z}_2$ degeneracy. In fact, the degeneracy can be related to the action of diffeomorphisms on $\mathcal{Q}$. First, however, I will discuss generically how this degeneracy comes about.
Start with the $q=0$ sector denoted $\mathcal{Q}_0$. In this sector, the configuration should asymptote to $\{\ou{0}{\mathcal{A}}{0}=- dh_t\,h_t^{-1},\gamma_5\}$. As in \cite{Giulini:1993gd}, to analyze the topological properties, it is convenient to define a canonical homeomorphism between each topological sector and a space where the analytic properties of of the map are more apparent. Thus, for each sector $\mathcal{Q}_q$, we define a homeomorphism to a new sector $\mathcal{Q}^*_0$, which is itself homotopic to $\mathcal{Q}_0$. The homeomorphisms is defined as follows (see Fig. \ref{Homeomorphism}). Given a configuration in the $\mathcal{Q}_q$ sector first transform to the EC gauge where $V=\gamma_5$. Next map the pair $\{\mathcal{A}, \gamma_5\}\rightarrow \{h_t^{-1}\mathcal{A} h_t -dh_t^{-1} h_t, \gamma_5\}$. This serves to remove the time dependence of the fiducial configurations $\{\ou{m}{\mathcal{A}}{n}, \gamma_5\}$.
Now transform to the trivial gauge. Finally transform the configuration by $\{0,V\}\rightarrow \{0,\ou{0}{h}{q} V \ou{0}{h}{q}{}^{-1}\}$. The point behind this homeomorphism is that it transforms each of the fiducial configurations $\{\ou{m}{\mathcal{A}}{n},\gamma_5\}$ defined in the previous section to the convenient base point $\{0,\gamma_5\}$. This will make it easier to determine the homotopy properties of the configuration.
The resulting configuration lives in a space $\mathcal{Q}_0^*$. This space is formed by the set of maps $V:\mathbb{R}\times \mathbb{S}^3 \rightarrow \mathcal{G}/\mathcal{H}$ that asymptote to $V=\gamma_5$ in the asymptotic past and future. Because of the latter restriction, one can add the endpoints $t=\{+\infty\}$ and $t=\{-\infty\}$ and compactify the domain to $\mathbb{S}^4$. Thus $V:\mathbb{S}^4 \rightarrow \mathcal{G}/\mathcal{H}$. The homotopy classes of maps of this type are characterized by $\pi_4(\mathcal{G}/\mathcal{H})=\mathbb{Z}_2$. Thus, via this homeomorphism, all states in $\mathcal{Q}$ can be classified into two groups: those configurations that when mapped to $\mathcal{Q}^*_0$ are homotopic to $\{0,\gamma_5\}$ and those that are homotopic to a generator of $\pi_4(\mathcal{G}/\mathcal{H})=\mathbb{Z}_2$ in $\mathcal{Q}^*_0$.
\begin{figure}
\begin{center}
\includegraphics[height=6.0cm]{Homeomorphism.jpg}
\end{center}
\caption{\label{Exchange} \label{Homeomorphism} Schematic of the homeomorphism between the $\mathcal{Q}_q$ sectors and the reference space $\mathcal{Q}^*_0$. On the left is the $\mathcal{Q}_{-1}$ sector containing the de Sitter solution (black point) and the twisted de Sitter solution (red point) which are mapped to different $\mathbb{Z}_2$ sectors of $\mathcal{Q}^*_0$. }
\end{figure}
\subsection{Rotational diffeomorphisms and the $\mathbb{Z}_2$ degeneracy}
Now let us consider the action of diffeomorphisms on $\mathcal{Q}$. To preserve the phase space, one should restrict attention to those that asymptote to an isomorphism in the asymptotic past and future. However, since an isomorphism is equivalent to a gauge transformation by a constant element of $Spin(4,1)$ and are therefore contained in the local gauge group, we will restrict to $\mbox{\it Diff}_0$, the set of diffeomorphisms that tend to the identity at future and past asymptotic infinity.
Consider first the $q=-1$ sector where the fiducial configuration $\{\ou{0}{\mathcal{A}}{1},\gamma_5\}$ represents ordinary de Sitter space. Define a one-parameter rotational diffeomorphism $\varphi$ by its action on the coordinates $\varphi(\{t,\chi, \theta, \phi\})=\{t,\chi,\theta,\phi'\equiv\phi-\phi_0(t)\}$ where $\phi_0(t)$ is a smooth function of $t$ only that varies from $0$ to $2\pi$ in the interval $[t_i,t_f]$, and is constant outside the interval. Assume also that the derivative $\partial_t\phi_0$ vanishes at $t_i$ and $t_f$. For definiteness, choose $t_i=-\ell$ and $t_f=\ell$. Explicitly, the diffeomorphism is given by $\varphi=\exp(-\phi_0\mathcal{L}_{\bar{\phi}})$ where $\bar{\phi}=\frac{\partial}{\partial\phi}$, and its action on de Sitter space is represented by a $2\pi$ twist. The metric transforms to (see Fig. \ref{TwisteddS} for visualization)
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\ou{0}{\bm{g}}{1}=-(1-\alpha)dt^2 +2\beta \,dt\,d\phi' +\cosh^2(t/\ell)\left(d\chi^2 +\sin^2\chi\left(d\theta^2+\sin^2\theta\, d\phi'{}^2\right)\right) \label{TwisteddSMetric}
\eeq
where $\alpha=\cosh^2(t/\ell)\,\sin^2\chi \,\sin^2\theta \,(\partial_t\phi'_0)^2$ and $\beta=\cosh^2(t/\ell)\,\sin^2\chi \,\sin^2\theta \,\partial_t\phi'_0$.
Consider now the action of of $\varphi$ on the configuration $\{\ou{0}{\mathcal{A}}{1}, \gamma_5\}$. The diffeomorphism leaves $V=\gamma_5$ fixed in this gauge, but transforms $\ou{0}{\mathcal{A}}{1}=-d\ou{0}{g}{1}\,\ou{0}{g}{1}{}^{-1}$ to $\varphi(\ou{0}{\mathcal{A}}{1})=-d\ou{0}{g}{1}{}'\,\ou{0}{g}{1}{}'^{-1}$ where
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\ou{0}{g}{1}{}'=h_t U^{-1} \ou{0}{h}{1} U \quad \mbox{with} \quad U(t)=\left[\begin{matrix} \exp(\phi_0 \,\frac{i}{2}\sigma_3) & 0 \\ 0 & \exp(\phi_0 \,\frac{i}{2}\sigma_3)\end{matrix}\right]
\eeq
Under the canonical homomorphism to $\mathcal{Q}^*_0$ given above, this configuration maps to
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
V^{\h{i}}=(u^{-1}h u h^{-1})^{\h{i}} \quad \mbox{with}\quad u= \exp\left(\phi_0 \,\textstyle{\frac{i}{2}}\sigma_3\right)
\eeq
where it is understood in this expression that the components $V^{\h{i}}$ are extracted from the expression $V=V^4 \bm{1}+V^i \,i\sigma_i$. I claim that this is a generator of $\pi_4(\mathcal{G}/\mathcal{H})=\mathbb{Z}_2$. To see this, I must first digress to discuss more generally the generators of $\pi_4(\mathbb{S}^3)=\mathbb{Z}_2$ (see e.g. \cite{Finkelstein:1968hy,Giulini:1993gd,Williams:Skyrme} for a related discussion in the context of the Skyrme model).
\subsection{The Hopf map and suspensions}
The group $\pi_4(\mathbb{S}^3)=\mathbb{Z}_2$ has just two elements, so it is sufficient to find a single map taking $\mathbb{S}^4\rightarrow \mathbb{S}^3$ that is not deformable to the identity. Any map that is homotopic to this map will then also be a generator of $\mathbb{Z}_2$. To construct the generator, consider first the lower dimensional case of maps from $\mathbb{S}^3$ onto $\mathbb{S}^2$. Since $\pi_3(\mathbb{S}^2)=\mathbb{Z}$, this group is also generated by a single map, and the canonical generator is known as the Hopf map. This map can be thought of as a fibration of the three-sphere into a non-trivial bundle of $U(1)$ fibers over $\mathbb{S}^2$. The projection map of the bundle $\pi:P=\mathbb{S}^3\rightarrow M=\mathbb{S}^2$ is the Hopf map itself. Explicitly it can be constructed as follows. Consider a vector field $n^{i}$ in the tangent space of $\mathbb{S}^3$ whose magnitude $\delta_{ij}n^i n^j=1$ constrains it to live in a $\mathbb{S}^2$ submanifold of the tangent space. The group $SU(2)$ acts transitively on this space via the adjoint action $n=n^{i}\,i\sigma_i \rightarrow n' =ana^{-1}$ for any $a:\mathbb{S}^3 \rightarrow SU(2)$. Now, take $a$ to be the standard generator, $h$, of $\pi_3(\mathbb{S}^3)$ given in (\ref{Zgenerator}), and take $n^i=(0,0,1)$. Then, since $h$ is not deformable to the identity and has winding number one, the map $n'=hnh^{-1}=h \,i\sigma_3 h^{-1}$ is a map from $\mathbb{S}^3$ to $\mathbb{S}^2$ with winding number one. It therefore serves as the generator of $\pi_3(\mathbb{S}^2)=\mathbb{Z}$. This is the Hopf map.
\begin{wrapfigure}{r}{0.4\textwidth}
\begin{center}
\includegraphics[width=0.4\textwidth]{TwisteddS.jpg}
\end{center}
\caption{\label{TwisteddS} The twisted version of de Sitter space given by the metric (\ref{TwisteddSMetric}). The red lines represent lines of constant $\phi'$.}
\end{wrapfigure}
A well known result of homotopy theory is that any suspension of the Hopf map is a generator of $\pi_4(\mathbb{S}^3)=\mathbb{Z}_2$. An example of such a suspension is any continuous map $V:\mathbb{S}^4\rightarrow \mathbb{S}^3$ such that the restriction of the map to the $\mathbb{S}^3$ equator of $\mathbb{S}^4$ reduces to the Hopf map. For example, take $\chi$ to be the azimuthal angle on $\mathbb{S}^4$. Then the map $V^{\h{i}}=(huh^{-1})^{\h{i}}$, where as before $u=\exp\left(2\chi \,\frac{i}{2}\sigma_3 \right)$, is such a suspension since on the equator at $\chi=\frac{\pi}{2}$, the map reduces to $V(\pi/2)=h\,i\sigma_3 h^{-1}$, the Hopf map.
\subsection{The $\mathbb{Z}_2$ configurations in $\mathcal{Q}$}
We now return to our configuration $V^{\h{i}}=(u^{-1}h u h^{-1})^{\h{i}}$. Evaluated at $t=0$, this map reduces to $-i \sigma_3 h\, i\sigma_3 h^{-1}$. This is simply the Hopf map followed by a constant rotation by $-\pi$, and is therefore homotopic to the Hopf map. In turn, the full map $V^{\h{i}}$ is homotopic to a suspension of the Hopf map in $\mathbb{S}^4$, and is therefore a generator of $\pi_4(\mathcal{G}/\mathcal{H})=\mathbb{Z}_2$. Thus, we have constructed the generator of the $\mathbb{Z}_2$ degeneracy in the sector $\mathcal{Q}_{-1}$. Clearly a similar construction holds in the sector $\mathcal{Q}_{1}$. Moreover, it should be clear that any time dependent spatial rotation of the same generic form as $\varphi$ is a generator of $\mathbb{Z}_2$. It is slightly less obvious, but still true, that a time dependent spatial {\it translation} of de Sitter space about one full revolution is also a generator\footnote{The easiest way to see this is to picture de Sitter space as the hyperboloid embedded in $\mathbb{R}^{1,4}$ with coordinates $\{T,X,Y,Z,W\}$. A global $2\pi$-rotation about the $XY$ plane is clearly homotopic to a global $2\pi$-rotation about $XZ$ since any spatial plane can be continuously deformed into another. By the same token it is homotopic to a $2\pi$-rotation about the $XW$ plane. But the latter is interpreted as a translation of the spacetime about one full revolution. The same reasoning applies for time dependent rotations and translations.} of $\mathbb{Z}_2$.
To construct the generator in all sectors $\mathcal{Q}_q$ we borrow well known results from the Skyrme model \cite{Finkelstein:1968hy,Giulini:1993gd,Williams:Skyrme}. Consider the action of $\varphi$ on the base points $\{\ou{m}{\mathcal{A}}{n},\gamma_5\}$, where we recall $q\equiv m-n$. It can be shown that in any sector $\mathcal{Q}_q$ with {\it odd} charge $q$, the action of $\varphi$ on the base point $\{\ou{m}{\mathcal{A}}{n},\gamma_5\}$ is a generator of $\mathbb{Z}_2$. On the other hand, the map is deformable to the trivial map in any sector with {\it even} $q$. In sectors of even charge, the generator can be constructed from exchanges of the geometries (see Fig. \ref{Exchange}).
\begin{figure}
\begin{center}
\includegraphics[height=6.0cm]{Exchange.jpg}
\end{center}
\caption{\label{Exchange} The exchange of two copies of de Sitter space in the $q=\pm 2$ sector. This map is homotopic to $2\pi$ twist of just one of the copies, and is therefore a generator of $\mathbb{Z}_2$ in the $q=\pm 2$ sector.
}
\end{figure}
Consider for example the $q=-2$ sector whose base point consists of two copies of de Sitter space attached along a degenerate surface. A diffeomorphism representing an exchange of these two copies is homotopic to a $2\pi$ twist of just one copy. Thus, the exchange is the generator of $\mathbb{Z}_2$ in this case. Similar results hold in all sectors with even $q$.
\section{Fermionic Quantization}
I now turn to the quantization of the generalized de Sitter spaces. Our goal here is to present the overall structure of a quantization scheme, with emphasis on the ways that it differs from a more conventional quantization. Most importantly, one must pay particularly careful attention to the role of symmetries, exact or asymptotic, and their effect on the wavefunction. As with asymptotically flat spacetimes, given our asymptotic (generalized) de Sitter past and future boundary conditions, one must make a distinction between an arbitrary diffeomorphism and a diffeomorphism representing an isometry. The latter are genuine symmetries of the system and the associated generators of the symmetries in the canonical theory are generally associated with physical quantities (e.g. energy, momentum, angular momentum, etc...). As a consequence, a diffeomorphism representing an isometry or asymptotic isometry should act non-trivially on the wavefunction in the quantum theory.
Since I have restricted attention to the space of flat connections, all our solutions are maximally symmetric solutions with ten Killing vectors. Choose two spatial hypersurfaces $\Sigma_1$ and $\Sigma_2$, for simplicity chosen to be located at finite values $t_1$ and $t_2$. Without loss of generality, assume $t_1 <0$ and $t_2>0$. Fix the geometry on the hypersurfaces to be the pull-back of (\ref{AConfig}) and (\ref{GConfig}) in the canonical coordinate system $\{\chi, \theta, \phi\}$. We wish to consider the set of diffeomorphisms that preserves the embedding of the spatial hypersurfaces $\Sigma_1$ and $\Sigma_2$ and their respective geometries. Naturally, this includes the set of isometries of $\ou{m}{\bm{g}}{n}$. So, let $\overline{Diff}$ be the set of identity connected diffeomorphisms that preserve the embedding of the hypersurfaces $\Sigma_1$ and $\Sigma_2$ {\it and} the pull-back of the metric to these surfaces, and let $\overline{Diff}_0 \subset \overline{Diff}$ be the subgroup that restricts to the identity on the boundary. The coset $\overline{Diff}/\overline{Diff}_0$ generically acts non-trivially on the wavefunction in the quantum theory.
To simplify the analysis, we will restrict attention to the $|q|=1$ de Sitter case and consider the action of $\overline{Diff}/\overline{Diff}_0$ on it. A typical element of the coset space is given by a pair $\{g_1, g_2\}$ representing the action of the diffeomorphism on the two endcaps $\Sigma_1$ and $\Sigma_2$, possibly subject to some as yet undefined equivalence relations. Given a constant-$t$ hypersurface of the manifold, the action of the subgroup of diffeomorphisms that preserve the hypersurface and the restriction of the metric to the hypersurface can be represented by a group $SO(4)_{\Sigma_t}$. Thus, $g_1\in SO(4)_{\Sigma_{t_1}}$ and $g_2\in SO(4)_{\Sigma_{t_2}}$.
Consider first a one-parameter family of diffeomorphisms $\varphi_\epsilon$ representing a smooth twist of the manifold as $\epsilon$ ranges from $0$ to $1$. Suppose that for each value $\epsilon$, the diffeomorphism represents a $2\pi\epsilon$ rotation in the $\frac{\partial}{\partial\phi}$ direction of the $\Sigma_2$ hypersurface while keeping $\Sigma_1$ fixed. In the interior between $t_i$ and $t_f$, the diffeomorphism smoothly interpolates between the identity on $\Sigma_1$ and the rotation on $\Sigma_2$. For example, the action of the diffeomorphism could be represented by its action on the coordinates by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\varphi_\epsilon \{\chi, \theta,\phi\} = \{\chi, \theta,\phi'\equiv\phi-\epsilon \phi_0(t)\}
\eeq
where $\phi_0$ is defined as before. Viewed as an element of $\overline{Diff}/\overline{Diff}_0$, the twist can be represented by the pair $\{g_1,g_2\}=\{1, g(\epsilon)\}$ where $g(\epsilon)$ represents a $2\pi\epsilon$ rotation in $SO(4)$. It can be shown that the action of the diffeomorphism on the boundary data alone is a loop parameterized by $\epsilon$ in the set of boundary data that is contractable if $q$ is even, and non-contractable if $q$ is odd. Thus, the fact that $\varphi_{\epsilon=1}$ is a generator of $\mathbb{Z}_2$ only in the odd sectors is matched by the pure boundary description.
In the boundary description appropriate for canonical quantization, one first identifies a configuration space $\mathcal{C}$ on each boundary by choosing a polarization of the phase space of boundary data equipped with a symplectic structure. Although the configuration space will depend on the polarization and the details of the quantization we will make the assumption that a $2\pi$ rotation is a non-contractable loop in the configuration space\footnote{If, for example, one were to consider the phase space corresponding to Einstein-Cartan theory via the EC action restricted to the set of flat Cartan connections, then this property would hold.}. Following alongside Skyrme theory, the formal procedure for fermionic quantization is to first take the double cover $\overline{\mathcal{C}}$ of the configuration space prior to quantization. In practice this is difficult to do, so fermionic quantization usually proceeds by first identifying the key non-contractable loops and promoting them to operators on the Hilbert space. The constraints, which go by the name of Finkelstein-Rubenstein constraints \cite{Finkelstein:1968hy} in Skyrme theory, associate a $-1$ phase to non-contractable loops. In other words consider a parameterized loop $\gamma:\mathbb{S}^1\rightarrow \mathcal{C}$, parameterized by $\epsilon \in[0,1]$, generated by a Hamiltonian flow. Promoting this to a parameterized set of operators $\h{\mathcal{O}}_{\gamma(\epsilon)}$ acting on the Hilbert space, we require
\beqa
\h{\mathcal{O}}_{\gamma(1)}|\Psi\rangle =\begin{cases} \ \ |\Psi\rangle & \mbox{if $\gamma(\epsilon)$ is contractable} \\ -|\Psi\rangle &\mbox{if $\gamma(\epsilon)$ is non-contractable} \end{cases} \,.
\end{eqnarray}}\newcommand{\w}{\wedge}\newcommand{\del}{\nabla}\newcommand{\ts}{\textstyle
In our case, the $\mathbb{Z}_2$ generators are given by the action of $\overline{Diff}/\overline{Diff}_0$ on the two endcaps of the manifold, represented by $\overline{\mathcal{C}}_{\Sigma_1} \cup \overline{\mathcal{C}}_{\Sigma_2}$. Restricting to the $q=\pm 1$ sector containing ordinary de Sitter space, we can easily guess the action of this group by considering the generators of the $\mathbb{Z}_2$ degeneracy on $\mathcal{Q}_{\pm 1}$. We first recall that $\overline{Diff}$ consists of the set of identity connected diffeomorphisms that restrict to hypersurface and geometry preserving diffeomorphisms on the boundary $\partial M=\Sigma_1 \cup \Sigma_2$. The subgroup of the isometry group of de Sitter space that preserves the embedding of a spatial hypersurface with topology $\mathbb{S}^3$ is $SO(4)$, which consists of the the spatial rotations and compact spatial translations. Taking the double cover, we expect that $\overline{Diff}/\overline{Diff}_0$ can be identified with some appropriate equivalence class of $Spin(4)_{\Sigma_1}\times Spin(4)_{\Sigma_2}$. To find this equivalence class, we consider three separate cases.
First, consider the action of a diffeomorphism representing a rotational or spatial translational isometry in a fixed direction. As an element of $\overline{Diff}/\overline{Diff}_0$, the action of the isometry on $\overline{\mathcal{C}}_{\Sigma_1} \cup \overline{\mathcal{C}}_{\Sigma_2}$ can be represented by the pair $\{g_1, g_2\}=\{g(\epsilon), g(\epsilon) \}$ where $\epsilon$ ranges from $0$ to $1$ and $g(\epsilon)$ represents a $2\pi\epsilon$ rotation. Such an isometry is {\it not} a generator of the $\mathbb{Z}_2$ degeneracy as can be seen as follows. The action of the diffeomorphism on $\mathcal{Q}$ can be constructed by fixing the the data at $t=0$ and twisting each endcap by $\epsilon$ while keeping $\Sigma_0$ fixed. At $\epsilon=1$, the lower and upper halves are both generators of $\mathbb{Z}_2$ on $\mathcal{Q}$, but the product of two generators of $\mathbb{Z}_2$ is itself deformable to the identity. As a consequence, we should expect that for an isometry, $\{g(0),g(0)\} \approx \{g(1),g(1)\}$. Since $g(1)=-1$, we have $\{1,1\}\approx\{-1,-1\}$.
Next consider a true generator of $\mathbb{Z}_2$ on $\mathcal{Q}$. This can be constructed in two equivalent ways. One way is to fix the hypersurface $\Sigma_1$ and rotate or translate only the other endcap $\Sigma_2$. This can be represented by the pair $\{g_1, g_2\}=\{1, g(\epsilon)\}$. The endpoint of the twist at $\epsilon=1$ is therefore $\{1,-1\}$. On the other hand, the generator can also be constructed by holding $\Sigma_2$ fixed and rotating or translating $\Sigma_1$ in the opposite direction. In this case the action is represented by the pair $\{g_1, g_2\}=\{g(\epsilon)^{-1}, 1\}$. The endpoint of the twist in this case is given by $\{-1,1\}$. However, since both cases represent the same action at the endpoint $\epsilon=1$, we must have $\{1,-1\}\approx \{-1,1\}$.
In total this implies that the action of $\overline{Diff}/\overline{Diff}_0$ on the configuration space is given by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\frac{Spin(4)\times Spin(4)}{\{-1,-1\}} \ \vartriangleright \ \overline{\mathcal{C}}_{\Sigma_1} \cup \overline{\mathcal{C}}_{\Sigma_2}\,.
\eeq
One consequence of this is that even when the double cover is taken, the action of the set of boundary preserving isometries on $\overline{\mathcal{C}}_{\Sigma_1} \cup \overline{\mathcal{C}}_{\Sigma_2}$ is still isomorphic to $SO(4)$.
\subsection{The fermionic inner product}
Since we now understand the action of $\overline{Diff}/\overline{Diff}_0$ on $\overline{\mathcal{C}}_{\Sigma_1} \cup \overline{\mathcal{C}}_{\Sigma_2}$ (in the $q=\pm 1$ sector) it will be easiest to consider the inner product between an initial state $|\Psi_1\rangle$ on $\Sigma_1$ and the final state $|\Psi_2 \rangle$ on $\Sigma_2$.
First, consider an inner product that carries a faithful representation of the full group $\overline{Diff}/\overline{Diff}_0$. Since the action of the isometry subgroup is isomorphic to $SO(4)$, the inner product can be thought of as a vector representation of $SO(4)$, hence, it carries an index denoted $\langle \Psi_2 | \Psi_1 \rangle {}^{\hat{I}}$. The absolute-square is then given by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
|\langle \Psi_2 | \Psi_1 \rangle|^2 \equiv \eta_{\hat{I}\hat{J}} \,\langle \Psi_2 | \Psi_1 \rangle {}^{\hat{I}} \,\langle \Psi_2 | \Psi_1 \rangle {}^{\hat{J}} \,.
\eeq
To be more specific, let us work a particular representation. First associate 4-component Dirac spinors (or some appropriate subset therein) $\psi_1$ and $\psi_2$ with the initial and final states. Defining $\gamma^{\hat{I}}=\{\gamma^I,\gamma_5\}$, the intermediate inner product might be given by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\langle \Psi_2 | \Psi_1 \rangle {}^{\hat{I}}\equiv \bar{\psi}_2 \gamma^{\h{I}} \psi_1 \,.
\eeq
This inner product is chosen because it is faithful to the action of $\overline{Diff}/\overline{Diff}_0$ on $\overline{\mathcal{C}}_{\Sigma_1} \cup \overline{\mathcal{C}}_{\Sigma_2}$. Note that the action of $\{g_1,g_2\}$ is given by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\langle g_2 \Psi_2 | g_1 \Psi_1 \rangle {}^{\hat{I}}\equiv \bar{\psi}_2 g_2^{-1} \gamma^{\h{I}} g_1 \psi_1 \,.
\eeq
The isometry subgroup therefore acts on the inner product by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\langle g \Psi_2 | g \Psi_1 \rangle {}^{\hat{I}}\equiv \bar{\psi}_2 g^{-1} \gamma^{\h{I}} g \psi_1 =g^{\h{I}}{}_{\h{J}}\,\bar{\psi}_2 \gamma^{\h{J}} \psi_1
\eeq
where $g^{\h{I}}{}_{\h{J}}$ is the adjoint representation of $Spin(4)$, which is itself isomorphic to $SO(4)$ as should be expected from an isometry. It follows that $|\langle \Psi_2 | \Psi_1 \rangle|^2$ is invariant under isomorphisms. However, under a twist corresponding to a generator of the $\mathbb{Z}_2$ degeneracy of $\mathcal{Q}$, the inner product transforms by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\langle \Psi_2 | \Psi_1 \rangle {}^{\hat{I}} \quad \longrightarrow \quad \langle g_2\Psi_2 | g_1\Psi_1 \rangle {}^{\hat{I}}=-\langle \Psi_2 | \Psi_1 \rangle {}^{\hat{I}}\,.
\eeq
Alternatively, we can define an inner product that is invariant under isomorphisms, but still retains information about twists by
\begin{equation}}\newcommand{\eeq}{\end{equation}}\newcommand{\beqa}{\begin{eqnarray}
\langle \langle \Psi_2 |\Psi_1 \rangle \rangle \equiv \bar{\psi}_2 \psi_1\,.
\eeq
\section{Concluding Remarks}
In ordinary quantum field theory, it is generally taken as given that bosons are comprised of ordinary commuting fields, and fermions are comprised of Grassman fields. However, the non-linear structure of certain field theories admits an alternative means for fermionic statistics to emerge. Here I have given strong evidence that this possibility may be realized in quantum gravity. Together with our previous work on emergent fermions in torsional systems \cite{Randono:2010cd,Randono:2011bb}, we now have two examples of fermionic geometries. Regarding the present work, the prospect that an entire spacetime could itself behave fermionically challenges many implicit tenets of quantum gravity, quantum cosmology, and even quantum field theory. Furthermore, the spacetime geometry in question is hardly exotic, being the simplest model spacetime beyond Minkowski space. Thus, the physical ramifications of fermionic geometries need to be explored in depth.
\section*{Acknowledgment}
This work was supported in part by the NSF International Research Fellowship grant OISE0853116.
\begin{appendix}
\section{Clifford algebra conventions}
Here I will review the basic conventions used in this paper regarding the Clifford algebra notation. The Clifford algebra is spanned by a set of Clifford matrices $\{\gamma^I\}$ satisfying $\gamma^I \gamma^J +\gamma^J\gamma^I=2\eta^{IJ}$ where $\eta^{IJ}=diag(-1,1,1,1)$. The volume element $\gamma_5$ is defined to be $\gamma_5=\frac{i}{4!}\epsilon_{IJKL}\gamma^I\gamma^J\gamma^K\gamma^L=i\gamma^0\gamma^1\gamma^2\gamma^3$, where it is understood that the alternating symbol $\epsilon_{IJKL}$ is such that $\epsilon_{0123}=-\epsilon^{0123}=1$.
The ten dimensional de Sitter algebra, $\mathfrak{spin}(4,1)$ is spanned by the basis elements $\{\frac{1}{2} \gamma^{[I}\gamma^{J]}\,,\, \frac{1}{2}\gamma_5 \gamma^K \}$. The Lorentz stabilizer subalgebra, denoted $\mathfrak{h}=\mathfrak{spin}(3,1)$ is spanned by $\{\frac{1}{2}\gamma^{[I}\gamma^{J]}\}$ and its complement $\mathfrak{p}$ in $\mathfrak{spin}(4,1)$ is spanned by $\{\frac{1}{2}\gamma_5 \gamma^K\}$. Given an antisymmetric Lorentz valued matrix $A^{IJ}=A^{[IJ]}$, the index free object $A$ is given by $A\equiv \frac{1}{4}\gamma_{[I}\gamma_{J]}\,A^{IJ}$.
The $\mathfrak{spin}(4,1)$ connection coefficient in a local trivialization denoted by $\mathcal{A}$ splits correspondingly. However, it is convenient to represent the tetrad as simply $e\equiv \frac{1}{2} \gamma_I \, e^I$ which pulls the $\gamma_5$ out of the decomposition $\mathcal{A}=\omega \oplus \frac{1}{\ell} \gamma_5 \,e$.
\end{appendix}
|
\section{ Introduction }
The quantum Ising model
\begin{eqnarray}
{\cal H} = - \sum_{<i,j>} J_{i,j} \sigma^z_i \sigma^z_j - \sum_i h_i \sigma^x_i
\label{hdes}
\end{eqnarray}
where the nearest-neighbor couplings $J_{i,j}>0$
and the transverse-fields $h_i>0$ are independent random variables
drawn with two distributions $\pi_{coupling}(J)$ and $\pi_{field}(h)$
is the basic model to study quantum phase transitions at zero-temperature
in the presence of frozen disorder.
In dimension $d=1$, exact results for a large number of observables
have been obtained by Daniel Fisher \cite{fisher}
via the asymptotically exact Strong Disorder renormalization procedure
(for a review, see \cite{review_strong}). In particular, the transition is governed
by an Infinite Disorder fixed point
and presents unconventional scaling laws with respect to the pure case.
In dimension $d>1$, the Strong Disorder renormalization procedure can still be defined.
It cannot be solved analytically, because the topology of the lattice changes upon renormalization,
but it has been studied numerically with the conclusion that the transition is also governed by
an Infinite Disorder fixed point in dimensions $d=2,3,4$
\cite{motrunich,fisherreview,lin,karevski,lin07,yu,kovacsstrip,
kovacs2d,kovacs3d,kovacsentropy,kovacsreview}.
These numerical renormalization results
are in agreement with the results of independent quantum Monte-Carlo
in $d=2$ \cite{pich,rieger}.
Even if it is clear that the most natural method to study Infinite Disorder fixed points
is the Strong Disorder renormalization approach, it seems useful to determine whether
other approaches are able to describe Infinite Disorder scaling.
In this paper, we introduce a simple non-linear transfer approximation
for the surface magnetization
in finite dimension $d>1$, which is inspired from the
'Cavity-Mean-Field' approximation
developed in Refs \cite{ioffe,feigelman,dimitrova},
and we study numerically the critical properties
of this approximation in dimensions $d=2$ and $d=3$,
The paper is organized as follows.
In Section \ref{sec_transfert}, we recall briefly the
'Cavity-Mean-Field' approximation
developed in Refs \cite{ioffe,feigelman,dimitrova}
and introduce the non-linear transfer approach for finite dimensions $d>1$.
Our numerical results in dimension $d=2$ and $d=3$ are presented in sections \ref{sec_dim2}
and \ref{sec_dim3} respectively.
Our conclusions are summarized in section \ref{sec_conclusion}.
\section{ Non-linear transfer approach for the surface magnetization }
\label{sec_transfert}
\subsection{ 'Cavity-Mean-Field' approximation on the Cayley tree \cite{ioffe,feigelman,dimitrova} }
For the random quantum Ising model model defined on a tree of coordinence $(K+1)$,
the following 'Cavity-Mean-Field' approximation has been developed \cite{ioffe,feigelman,dimitrova} :
an ancestor $i$ is submitted to the effective single spin Hamiltonian
\begin{eqnarray}
H_i^{eff} = - B_i \sigma^z_i - h_i \sigma^x_i
\label{heffbihi}
\end{eqnarray}
where $h_i$ is its own random transverse field, and where $B_i$
represents the longitudinal field created by the $K$ children $j$ (related to $i$ by the
ferromagnetic couplings $J_{ij}$)
within a 'Mean-Field approximation' ( the operator $\sigma^z_j $ is replaced by its expectation value
$< \sigma^z_j > $)
\begin{eqnarray}
B_i=\sum_{j=1}^K J_{i,j} < \sigma^z_j >
\label{bimf}
\end{eqnarray}
The effective Hamiltonian of Eq. \ref{heffbihi} is only a two-level system
that can be solved immediately : the magnetization of the ground state reads
\begin{eqnarray}
m_i \equiv <\sigma_i^z>_{H_i^{eff} } = \frac{B_i}{\sqrt{B_i^2+h_i^2}}
\label{miheff}
\end{eqnarray}
Using Eq. \ref{bimf}, one obtains the following non-linear recurrence
for the magnetizations $m_i$
(see Eq. 4 of \cite{ioffe}, Eq. 7 of \cite{feigelman}, Eq. 17
of \cite{dimitrova} in the limit of zero temperature $\beta=+\infty$)
\begin{eqnarray}
m_i = \frac{ \sum_{j=1}^K J_{i,j} m_j }{\sqrt{ \left( \sum_{j=1}^K J_{i,j} m_j \right)^2+h_i^2}}
\label{mimj}
\end{eqnarray}
We refer to Refs \cite{ioffe,feigelman,dimitrova} for more details on
this 'Cavity-Mean-Field'
approximation and on its properties.
As a final remark, let us stress that the 'Cavity-Mean-Field' is not exact for the pure model
on the Cayley tree (see Fig. 3 of Ref \cite{dimitrova}), but has been argued to become
quantitatively correct in the limit of high connectivity $K \gg 1$ \cite{ioffe,feigelman,dimitrova}.
In the disordered phase where the magnetizations flows towards zero,
the non-linear recurrence of Eq. \ref{mimj} can be linearized to give the following
recursion
\begin{eqnarray}
m_i \simeq \frac{1}{h_i} \sum_{j=1}^K J_{i,j} m_j
\label{mimjlinear}
\end{eqnarray}
which is equivalent to the problem of a Directed Polymer on the Cayley tree
\cite{ioffe,feigelman,dimitrova}. This equivalence can be justified directly
at the level of lowest-order perturbation theory
(i.e. without invoking the 'Cavity-Mean-Field' approximation of Eq. \ref{mimj}),
and can be in this way extended to the finite-dimensional case \cite{transverseDP}.
\subsection{ 'Cavity-Mean-Field' approximation in $d=1$ \cite{dimitrova} }
\label{Cavity1d}
For $K=1$, the Cayley tree of coordinence $(K+1)$ discussed
in the previous section
becomes a one-dimensional chain, and Eq. \ref{mimj} becomes the
one-dimensional non-linear
recurrence \cite{dimitrova}
\begin{eqnarray}
m_i = \frac{ J_{i,i+1} m_{i+1} }{\sqrt{ \left( J_{i,i+1} m_{i+1} \right)^2+h_i^2}}
\label{mimjd1}
\end{eqnarray}
Assuming one starts with the boundary condition $m_L=1$ at site $i=L$,
one obtains
the following explicit expression for the surface magnetization
$m_0^{surf}$ at the site $i=0$
\cite{dimitrova}
\begin{eqnarray}
m_0^{surf}=
\left[ 1+ \sum_{i=0}^{L-1} \prod_{j=0}^i \left( \frac{h_j}{J_{j,j+1} } \right)^2 \right]^{-1/2}
\label{msurfexact}
\end{eqnarray}
As stressed in \cite{dimitrova}, this expression exactly coincides
with the rigorous expression
that can be obtained from a free-fermion representation \cite{peschel,msurf},
and from which many critical exponents can be obtained
\cite{msurf,dhar,ckesten}.
The reason why the 'Cavity-Mean-Field' approximation turns out to become
exact for the surface magnetization in $d=1$ is not clear to us, and seems rather surprising :
usually 'mean-field approximation' are exact in sufficiently high dimensions or on trees,
and are not exact in low dimensions,
the 'worst case' being precisely $d=1$.
Here we have exactly the opposite conclusion :
the 'Cavity-Mean-Field' is not exact for the pure model on the tree (for the disordered case,
it is not known), but turns out to be exact in $d=1$, both for the pure and the disordered case.
In the absence of any satisfactory explanation for this unusual situation,
we tend to think that the exactness of the 'Cavity-Mean-Field' in $d=1$
is likely to be a 'coincidence' specific to this particular case,
from which one cannot draw general conclusions for the validity of this approach in higher dimensions
or for other quantum disordered models.
\subsection{ Non-linear transfer approach in finite dimension $d>1$ }
\subsubsection{ Description }
\begin{figure}[htbp]
\includegraphics[height=8cm]{geom2d.eps}
\caption{ Notations to define the non-linear transfer approach in $d=2$ :
we impose the boundary conditions $m(x=0,y)=1$ on the left boundary,
and we study the surface magnetizations $m(x=L,y)=1$
on the right boundary (see text for more details). }
\label{figgeom2d}
\end{figure}
In finite dimensions $d>1$, the authors of Ref \cite{dimitrova}
have proposed to use the 'Cavity-Mean-Field' approximation on a Cayley tree
with parameter $K=2d-1$ (to reproduce the connectivity of each spin).
In the present paper, we propose instead to extend Eqs \ref{mimj}
towards an appropriate non-linear transfer approach
for the surface magnetizations of a finite sample of volume $L^d$.
For clarity, let us first explain the procedure for the case $d=2$.
As shown on Fig. \ref{figgeom2d}, we consider a lattice containing $L^2$
spins : when $x$ is even ($x=0,2,4,..$), the coordinate $y$ takes the $L$
integer-values $y=1,2,...,L$ ;
when $x$ is odd ($x=1,3,...$), the coordinate $y$ takes the $L$
half-integer-values $y=3/2,5/2,...,L+1/2$.
The boundary conditions are periodic in $y$ with $y+L \equiv y$.
At $x=0$, we impose the boundary condition of unity magnetization
\begin{eqnarray}
m (x=0,y) = 1
\label{mix0}
\end{eqnarray}
and we are interested in the $L$ surface magnetizations $m(x=L,y)$
at the opposite boundary $x=L$.
For this situation, we propose to use the ideas of Eqs \ref{mimj}
within the following transfer approach.
We assume that we have already found the surface magnetizations
on the column $m(x-1,y)$, and we add another column of $L$ sites at $x$.
From the Cavity point of view, the new spin at $(x,y)$ is submitted
to its own random transverse field $h(x,y)$ and
to the longitudinal field (see Eq. \ref{bimf})
created by its two neighbors on the column $x-1$
\begin{eqnarray}
B(x,y)=
J_{\{\left(x,y\right), \left(x-1,y+\frac{1}{2}\right)\}}
m \left(x-1,y+\frac{1}{2}\right)
+J_{\{\left(x,y\right), \left(x-1,y-\frac{1}{2}\right)\}}
m \left(x-1,y-\frac{1}{2}\right)
\label{bimf2d}
\end{eqnarray}
so that its surface magnetization reads (Eq. \ref{miheff})
\begin{eqnarray}
m (x,y) = \frac{B(x,y)}{\sqrt{B^2(x,y)+h^2(x,y)}}
\label{miheff2d}
\end{eqnarray}
Eqs \ref{bimf2d} and \ref{miheff2d} define a non-linear transfer procedure
that can be iterated from the boundary condition on the column $x=0$
of Eq. \ref{mix0} up to $x=L$, where we analyze the statistics
of the final surface magnetizations $m(x=L,y)$.
It is clear that the generalization of this procedure
to $d=3$ is straightforward : we add another direction $z$
with periodic boundary conditions that plays exactly the same role as $y$.
\subsubsection{ Linearized transfer matrix within the disordered phase }
Within the disordered phase, the surface magnetizations $m(x=L,y)$
are expected to decay typically exponentially in $L$, so that
one may linearize the transfer Eqs \ref{bimf2d} and \ref{miheff2d}
to obtain
\begin{eqnarray}
{ \bf Linearization : \ \ } m (x,y) \simeq
\frac{J_{\{\left(x,y\right), \left(x-1,y+\frac{1}{2}\right)\}}}{h(x,y)}
m \left(x-1,y+\frac{1}{2}\right)
+ \frac{J_{\{\left(x,y\right), \left(x-1,y-\frac{1}{2}\right)\}}}{h(x,y)}
m \left(x-1,y-\frac{1}{2}\right)
\label{milinear}
\end{eqnarray}
This linearized equations can be derived directly within a lowest-order
perturbative approach \cite{transverseDP}
(i.e. without invoking the 'Cavity-Mean-Field'
approximation) and corresponds
to the transfer matrix satisfied by the
partition function of a Directed Polymer
with $D=(d-1)$ transverse directions,
as discussed in detail in \cite{transverseDP}.
We refer to \cite{transverseDP} for the description of the consequences
of this correspondence, and for the analogy with Anderson localization, where
the droplet exponent of the Directed Polymer also appears in the localized phase
\cite{NSS,medina,prior}.
Here our conclusion is that the
non-linear transfer approach describes at least correctly
the disordered phase, where it coincides with the lowest-order
perturbative approach \cite{transverseDP}.
\subsubsection{ Discussion }
Besides its correctness in the disordered phase that we have
just discussed, the validity of the non-linear transfer
exactly at criticality and in the ordered phase has to be studied
for the disordered case in $d>1$.
Since it has been found to be exact in $d=1$ (see section \ref{Cavity1d}),
one could hope that it is not 'too bad' in $d=2,3$
(even if it is clear that this approximation is not valid for the pure model) :
we believe that it should capture correctly the nature of the transition
between 'Infinite-Disorder' or 'Conventional' scaling.
In the following, we present our numerical results in $d=2$ and $d=3$
and discuss the scaling properties in the two phases and at criticality.
\section{ Numerical results in dimension $d=2$ }
\label{sec_dim2}
In this section, we present the numerical results obtained
with the following sizes $L$ and the corresponding numbers $n_s(L)$ of disordered samples
of volume $L^2$
\begin{eqnarray}
L && = 10^3, 2. 10^3 , 3.10^3 , 4.10^3 , 5.10^3 , 6.10^3, 7.10^3, 8. 10^3 \nonumber \\
n_s(L) && = 2.10^5, 13.10^4, 65.10^3, 37.10^3, 24.10^3, 17.10^3, 13.10^3, 10^4
\label{numed2}
\end{eqnarray}
For each sample $\alpha$, we collect the $L$ values of the surface
magnetization $m^{(\alpha)}(x=L,i)$
at the different points $i=1,2,..,L$ of the surface (see Fig. \ref{figgeom2d}).
Average values and histograms are then based on these $L \times n_s(L)$ values.
We have chosen to consider the following log-normal distribution
for the random transverse fields $h_i>0$
\begin{eqnarray}
\pi_{LN}(h) = \frac{1}{ h \sqrt{2 \pi \sigma^2 }} e^{- \frac{(\ln h - \overline{\ln h})^2}{2 \sigma^2 }}
\label{lognormal}
\end{eqnarray}
of parameter $ \overline{\ln h}=0$ and $\sigma=1$, whereas
the ferromagnetic couplings $J_{i,j}$ are not random
but take a single value $J$ that will be the control parameter of the quantum transition.
\subsection{ Disordered phase ($J<J_c$)}
\subsubsection{ Exponential decay of the typical surface magnetization
$m_L^{typ} \equiv e^{\overline{ \ln m_L^{surf} }}$ }
\begin{figure}[htbp]
\includegraphics[height=6cm]{xm24.2dall.eps}
\hspace{2cm}
\includegraphics[height=6cm]{xmloglog25.2daj=0.29.eps}
\caption{ Disordered phase $J<J_c$ in $d=2$
(a) Exponential decay of the typical surface magnetization
$m_L^{typ} \equiv e^{\overline{ \ln m_L^{surf} }}$ :
we show the linear decay of $ \ln m_L^{typ} $
as a function of the length $L$ (see Eq. \ref{msdis}).
(b) Log-log plot of the width $\Delta_L$
of the distribution of the logarithm of the surface magnetization
as a function of $L$ (here for $J=0.29$) : the slope is of order $\omega \simeq 0.33$
(see Eq. \ref{deltaLomega}). }
\label{fig2ddisordered}
\end{figure}
In the disordered phase $J<J_c$, one expects that the typical surface magnetization
defined by
\begin{eqnarray}
\ln (m_L^{typ}) \equiv \overline{ \ln m_L^{surf} }
\label{defmstyp}
\end{eqnarray}
decays exponentially
with $L$
\begin{eqnarray}
\ln (m_L^{typ}) \equiv \overline{ \ln m_L^{surf} }(J<J_c)
\mathop{\simeq}_{L \to \infty} - \frac{L}{\xi_{typ}(J)}
\label{msdis}
\end{eqnarray}
where $\xi_{typ}$ represents the typical correlation length that diverges
at the transition as a power-law
\begin{eqnarray}
\xi_{typ}(J) \mathop{\simeq}_{J \to J_c^-} (J_c-J)^{-\nu_{typ}}
\label{xityp}
\end{eqnarray}
On Fig. \ref{fig2ddisordered} (a) we show our numerical results :
concerning the exponential decay with $L$ of Eq. \ref{msdis}
for various values of $J$.
We find that the corresponding slope $1/\xi_{typ}(J)$
vanishes near the critical value $J_c \simeq 0.335$ with the exponent
\begin{eqnarray}
\nu_{typ} \simeq 1
\label{nutyp2d}
\end{eqnarray}
\subsubsection{ Growth of the width of the distribution of the logarithm of the surface magnetization }
In the disordered phase $J<J_c$, one expects that the width $\Delta_L$
of the distribution of the logarithm of the surface magnetization defined by
\begin{eqnarray}
\Delta_L \equiv \left( \overline{ (\ln m_L^{surf})^2 } - (\overline{ \ln m_L^{surf} })^2 \right)^{1/2}
\label{defdeltaL}
\end{eqnarray}
grows as a power-law of $L$
\begin{eqnarray}
\Delta_L (J<J_c) \mathop{\simeq}_{L \to \infty} L^{\omega}
\label{deltaLomega}
\end{eqnarray}
Our numerical data shown on Fig. \ref{fig2ddisordered} (b)
correspond to the value
\begin{eqnarray}
\omega(d=2) \simeq 0.33
\label{omega2d}
\end{eqnarray}
in agreement with the argument presented in \cite{transverseDP}
where $\omega(d=2)$ should coincide with the Directed Polymer droplet exponent
$\omega_{DP}(D=d-1=1)=1/3$ \cite{Hus_Hen_Fis,Kar,Joh,Pra_Spo}.
\subsubsection{ Distribution of the logarithm of surface magnetization }
\begin{figure}[htbp]
\includegraphics[height=6cm]{xm70.2daj=0.29.eps}
\hspace{1cm}
\includegraphics[height=6cm]{xm73.2daj=0.29.eps}
\caption{ Disordered phase in $d=2$ (here $J=0.29$) :
(a) Evolution with $L$ of the probability distribution $P_L(\ln m_L^{surf})$
of the logarithm of surface magnetization :
(b) Corresponding fixed distribution of the rescaled variable
$u=(\ln m_L^{surf} -\ln m_L^{typ})/\Delta_L $ in log-scale to show the tails,
compared to the exact Tracy-Widom GOE distribution (thick line).
}
\label{fighistodisordered}
\end{figure}
We show on Fig \ref{fighistodisordered} (a) our numerical results
concerning histograms of the logarithm of the surface magnetization
in the disordered phase. Our conclusion is that the surface magnetization follows the scaling
\begin{eqnarray}
\ln (m_L^{surf}) \mathop{\simeq}_{L \to \infty} \ln (m_L^{typ}) +\Delta_L u
\label{msdisfull}
\end{eqnarray}
where the behaviors of the typical value
$\ln (m_L^{typ}) \simeq -L/\xi_{typ}$
and of the width $\Delta_L \sim L^{\omega}$ have been already discussed
above in Eqs \ref{msdis} and \ref{deltaLomega} respectively.
On Fig. \ref{fighistodisordered} (b), we show that the stable distribution $P(u)$
of the rescaled variable $u$ coincides with the GOE
Tracy-Widom distribution, as expected from the correspondence with
the Directed Polymer model in the disordered phase \cite{transverseDP}.
\subsection{ Ordered phase }
\subsubsection{ Behavior of the typical surface magnetization $m_{\infty}^{typ}$
in the ordered phase $J>J_c$ }
\begin{figure}[htbp]
\includegraphics[height=6cm]{xm24limit2d.eps}
\hspace{1cm}
\includegraphics[height=6cm]{xm70.2daj=0.38.eps}
\caption{ Ordered Phase $J>J_c$ in $d=2$ :
(a)
Behavior of the asymptotic typical surface magnetization $m_{\infty}^{typ}$
as a function of the ferromagnetic coupling $J$ : our numerical data are
compatible with an essential singularity (Eq. \ref{defkappa})
of exponent $\kappa \simeq 0.5 $
(b)
the probability distribution $P_L(\ln m_L^{surf})$ (here for $J=0.38$)
of the logarithm of surface magnetization remains fixed and attached at the origin
(as a consequence of the bound $m_L^{surf} \leq 1$).
}
\label{fig2dordered}
\end{figure}
In the ordered phase, the typical surface magnetization remains finite
in the limit where the number of generations $L$ diverges
\begin{eqnarray}
\ln m_L^{typ}(J>J_c) \equiv \overline{ \ln m_L^{surf}(J>J_c) } \mathop{\simeq}_{L \to \infty}
\ln m_{\infty}(J>J_c) > -\infty
\label{transdeloc}
\end{eqnarray}
and one expects an
essential singularity behavior
\begin{eqnarray}
\ln m_{\infty}^{typ}(J>J_c) \oppropto_{J \to J_c^+} - (J-J_c)^{- \kappa}
\label{defkappa}
\end{eqnarray}
Our data shown on Fig. \ref{fig2dordered} (a)
can be fitted with the value
\begin{eqnarray}
\kappa (d=2) \simeq 0.5
\label{kappa2d}
\end{eqnarray}
that can be related to other exponents via finite-size scaling
(see below around Eq. \ref{kappaomega})
\subsubsection{ Distribution of the logarithm of the surface magnetization }
In the ordered phase, the probability distribution $P_L(\ln m_L^{surf})$
of the logarithm of surface magnetization remains fixed as $L$ varies,
and terminates discontinuously at the origin,
as a consequence of the bound $m_L^{surf} \leq 1$
corresponding to $\ln m_L^{surf} \leq 0$ (see Fig. \ref{fig2dordered} (b))
\subsection{ Critical point }
\subsubsection{ Behavior of the typical surface magnetization $m_{L}^{typ}$ and of the width $\Delta_L$ }
\begin{figure}[htbp]
\includegraphics[height=6cm]{xmloglog24et25.2daj=0.335.eps}
\hspace{1cm}
\includegraphics[height=6cm]{xmfss24d2.eps}
\caption{ Critical point (here $J_c=0.335$) :
(a) Log-log plot
of the logarithm of the typical surface magnetization $m_{L}^{typ}$
and of the width $\Delta_L$ : both slopes are of order $\omega_c \simeq 0.33$
(see Eqs \ref{mscriti} and \ref{deltaLcriti})
(b) Finite-size scaling of the typical surface magnetization
according to Eq. \ref{fssmtyp}
with $J_c=0.335$, $\omega_c=0.33$ and $\nu_{av}=1.5$. }
\label{fig2dcriti}
\end{figure}
Exactly at criticality, one expects that the typical surface magnetization
follows an activated behavior of exponent $\omega_c<1$ (compare with Eq. \ref{msdis} in the disordered phase)
\begin{eqnarray}
\ln (m_L^{typ}(J=J_c)) \equiv \overline{ \ln m_L^{surf}(J=J_c) }
\mathop{\simeq}_{L \to \infty} - L^{\omega_c}
\label{mscriti}
\end{eqnarray}
and that the width defined in Eq. \ref{defdeltaL} is also governed by the same exponent
\begin{eqnarray}
\Delta_L(J=J_c) \mathop{\simeq}_{L \to \infty} L^{\omega_c}
\label{deltaLcriti}
\end{eqnarray}
Our numerical data at $J_c \simeq 0.335$ shown on Fig. \ref{fig2dcriti} (a)
are compatible with these behaviors with the value
\begin{eqnarray}
\omega_c \simeq 0.33
\label{psicriti}
\end{eqnarray}
i.e. $\omega_c$ coincides with the fluctuation exponent $\omega$
measured in the disordered phase (see Eq. \ref{omega2d})
This last property implies that the finite-size scaling for the typical
surface magnetization $m_{L}^{typ}$ involves some correlation length
exponent $\nu_{av}$ different from $\nu_{typ}$
\begin{eqnarray}
\ln m_L^{typ}(J) \equiv \overline{ \ln m_L^{surf}(J)} \mathop{\simeq} - L^{\omega_c} G \left(x \equiv (J-J_c) L^{1/\nu_{av}} \right)
\label{fssmtyp}
\end{eqnarray}
The matching with the behavior of Eq. \ref{msdisfull}
in the disordered phase implies that
\begin{eqnarray}
G(x) \oppropto_{x \to -\infty} (-x)^{\nu_{typ}}
\label{Gxneg}
\end{eqnarray}
and that $\nu_{av}$ reads
\begin{eqnarray}
\nu_{av} = \frac{\nu_{typ}}{1-\omega}
\label{nuavnutyp}
\end{eqnarray}
This relation can be understood within a rare events analysis
for the averaged correlation in the disordered phase
\cite{transverseDP}.
The values $\nu_{typ} \simeq 1$ and $\omega=1/3$ yield
\begin{eqnarray}
\nu_{av}(d=2) \simeq \frac{3}{2}
\label{nuav2d}
\end{eqnarray}
The matching of the finite-size scaling form of Eq. \ref{fssmtyp}
with the essential singularity of Eq. \ref{defkappa}
in the ordered phase implies that
\begin{eqnarray}
G(x) \oppropto_{x \to + \infty} \frac{1}{x^{\kappa}}
\label{Gxpos}
\end{eqnarray}
with
\begin{eqnarray}
\kappa=\omega_c \nu_{av} = \nu_{typ} \frac{\omega}{1-\omega}
\label{kappaomega}
\end{eqnarray}
The values $\nu_{typ}=1$ and $\omega=1/3$ yield
\begin{eqnarray}
\kappa(d=2) \simeq \frac{1}{2}
\label{kappa2dtheo}
\end{eqnarray}
in agreement with the estimate of Eq. \ref{kappa2d}.
As shown on Fig \ref{fig2dcriti} (b), our numerical data collapse well
with the finite-size scaling form of Eq. \ref{fssmtyp} with $\nu_{av}=1.5$.
\subsubsection{ Distribution of the logarithm of surface magnetization }
\begin{figure}[htbp]
\includegraphics[height=6cm]{xm70insert.2daj=0.335.eps}
\hspace{1cm}
\caption{ Critical point in $d=2$ (here $J_c=0.335$) :
Evolution with $L$ of the probability distribution $P_L(\ln m_L^{surf})$
of the logarithm of surface magnetization.
Inset : Corresponding fixed distribution of the rescaled variable
$v=(\ln m_L^{surf})/\ln m_L^{typ} $.
}
\label{fighistocriti}
\end{figure}
At criticality, the rescaled variable
\begin{eqnarray}
v \equiv \frac{\ln m_L^{surf}}{\ln m_L^{typ}}
\propto - \frac{\ln m_L^{surf}}{L^{\omega_c}}
\label{mscritihisto}
\end{eqnarray}
remains a positive random variable of order $O(1)$ as $L \to +\infty$.
Our numerical measure of its probability distribution $P(v)$ shown on Fig. \ref{fighistocriti}
is compatible with a power-law singularity near the origin
\begin{eqnarray}
P(v) \mathop{\simeq}_{v \to 0^+} v^a
\label{pvcriti}
\end{eqnarray}
with an exponent of order $a \geq 2$ that we do not measure precisely.
Note that this is different from the case $d=1$ where $P(v=0)$ is finite ($a=0$).
We have not been able to find a physical argument to predict the value of $a$ in $d=2$.
This exponent $a$ will directly influence the scaling of
the moments of the surface magnetization, as we now discuss.
\subsubsection{ Moments of the surface magnetization }
In contrast to the activated behavior
of the typical surface magnetization $m_{L}^{typ}$
of Eq. \ref{mscriti}, the moments of the surface magnetization
are expected to follow a power-law, as a consequence of
the following rare events analysis :
the surface magnetization of Eq. \ref{mscritihisto}
will be of order $O(1)$ if the random variable $v$ happens
to be smaller than $1/L^{\omega_c}$. Taking into account the behavior
of Eq. \ref{pvcriti}, this will happen with probability
\begin{eqnarray}
Prob(m_L^{surf}=1) \simeq \int_0^{1/L^{\omega_c}} dv P(v) \sim \int_0^{1/L^{\omega_c}} dv v^a
\oppropto_{L \to \infty} L^{- \omega_c (1+a) }
\label{prob1xs}
\end{eqnarray}
and all moments will be governed by this power-law
\begin{eqnarray}
\overline{(m_L^{surf})^k} \simeq Prob(m_L^{surf}=1)
\oppropto_{L \to \infty} L^{- x_s } \ \ {\rm with } \ \ x_s=\omega_c (1+a)
\label{msavcriti}
\end{eqnarray}
independently of the order $k$.
Our numerical data for the three first moments $k=1,2,3$
and various sizes are compatible with Eq. \ref{msavcriti}
with an exponent
\begin{eqnarray}
x_s (d=2) \simeq 1.2
\label{xs2d}
\end{eqnarray}
The relation of Eq. \ref{msavcriti} then corresponds to
\begin{eqnarray}
a(d=2) \simeq 2.6
\label{a2d}
\end{eqnarray}
In the ordered phase, our numerical data are compatible with the power-law
\begin{eqnarray}
\overline{ m_L^{surf} } \propto (J-J_c)^{\beta_s}
\label{defbetas}
\end{eqnarray}
with
\begin{eqnarray}
\beta_s (d=2) = x_s \nu_{av} \simeq 1.8
\label{beta2d}
\end{eqnarray}
\section{ Numerical results in dimension $d=3$ }
\label{sec_dim3}
In this section, we present the numerical results obtained
with the following sizes $L$ and the corresponding numbers $n_s(L)$ of disordered samples
of volume $L^3$
\begin{eqnarray}
L && = 10^2, 2. 10^2 , 3.10^2 , 4.10^2 , 5.10^2 , 6.10^2, 7.10^2, 8. 10^2 \nonumber \\
n_s(L) && = 27.10^3, 7.10^3, 3.10^3, 16.10^2, 10^3, 7.10^2, 5.10^2, 4.10^2
\label{numed3}
\end{eqnarray}
For each sample $\alpha$, we collect the $L^2$ values of the surface
magnetization
at the different points of the surface.
Average values and histograms are then based on these $L^2 \times n_s(L)$ values.
We consider again the disorder distribution of Eq. \ref{lognormal}
and take $J$ as the control parameter of the transition.
\subsection{ Disordered phase }
\begin{figure}[htbp]
\includegraphics[height=6cm]{xm70.3daj=0.11.eps}
\hspace{1cm}
\includegraphics[height=6cm]{xm73.3daj=0.11.eps}
\caption{ Disordered phase in $d=3$ (here $J=0.11$) :
(a) Evolution with $L$ of the probability distribution $P_L(\ln m_L^{surf})$
of the logarithm of surface magnetization :
(b) Corresponding fixed distribution of the rescaled variable
$u=(\ln m_L^{surf} -\ln m_L^{typ})/\Delta_L $ in log-scale to show the
tails.
}
\label{fighistodisordered3d}
\end{figure}
Our data follow the scaling of Eq. \ref{msdisfull},
with the following properties :
(i) the scaling of the typical surface magnetization is given by
Eq \ref{msdis}, and the typical correlation length exponent of Eq. \ref{xityp},
seems again very close to unity
\begin{eqnarray}
\nu_{typ} \simeq 1
\label{nutyp3d}
\end{eqnarray}
(ii) the width $\Delta_L$ of Eq. \ref{defdeltaL} grows as the power-law
of Eq. \ref{deltaLomega}
with the exponent
\begin{eqnarray}
\omega(d=3) \simeq 0.24
\label{omega3d}
\end{eqnarray}
that coincides with the numerical values of
the droplet exponent of the Directed Polymer model with $D=d-1=2$
transverse dimensions \cite{Tan_For_Wol,Ala_etal,perlsman,KimetAla,Mar_etal,DPtails,schwartz},
in agreement with the argument presented in \cite{transverseDP}.
(iii) As $L$ grows, the evolution of the probability distribution $P_L(\ln m_L^{surf})$
is shown on Fig. \ref{fighistodisordered3d} (a).
The corresponding fixed distribution of the rescaled random variable
$u= (\ln m_L^{surf} -\ln m_L^{typ})/\Delta_L $ is shown
on Fig \ref{fighistodisordered3d} (b).
\subsection{ Critical point }
\begin{figure}[htbp]
\includegraphics[height=6cm]{xmloglog24et25.3daj=0.1528.eps}
\hspace{1cm}
\includegraphics[height=6cm]{xmfss24.3d.eps}
\caption{ Typical surface magnetization in the critical region in $d=3$
(a) Log-log plot of the logarithm of the typical surface magnetization $m_{L}^{typ}$
and of the width $\Delta_L$ at criticality $J_c=0.1528$ :
both slopes are of order $\omega_c \simeq 0.24$ (see Eqs \ref{mscriti} and \ref{deltaLcriti} )
(b) Finite-size scaling of the typical surface magnetization in $d=3$
according to Eq. \ref{fssmtyp} with $\nu_{av}=1.32$.
}
\label{figfss3d}
\end{figure}
At criticality $J_c=0.1528$, we find that the exponent $\omega_c$ of Eqs \ref{mscriti}
and \ref{deltaLcriti} coincides with the value of $\omega$ of Eq. \ref{omega3d}
concerning the disordered phase (see Fig. \ref{figfss3d} (a))
\begin{eqnarray}
\omega_c(d=3) \simeq 0.24
\label{omegas3d}
\end{eqnarray}
We show on Fig. \ref{figfss3d} (b) the finite-size scaling analysis
of Eq. \ref {fssmtyp} for the logarithm of the typical surface magnetization
with the averaged correlation length exponent $\nu_{av}=1/(1-\omega) \simeq 1.32$.
\begin{figure}[htbp]
\includegraphics[height=6cm]{xm70insert.3daj=0.1528.eps}
\hspace{1cm}
\caption{ Critical point in $d=3$ (here $J_c=0.1528$) :
Evolution with $L$ of the probability distribution $P_L(\ln m_L^{surf})$
of the logarithm of surface magnetization.
Inset : Corresponding fixed distribution of the rescaled variable
$v=(\ln m_L^{surf})/\ln m_L^{typ} $.
}
\label{fighistocriti3d}
\end{figure}
We show on Fig. \ref{fighistocriti3d}
our numerical data for the probability distribution of the surface magnetization
at criticality~: the fixed point distribution $P(v)$
of the rescaled variable $v$ of Eq. \ref{mscritihisto}
displays a power-law singularity near the origin
(Eq \ref{pvcriti}), that will determine the scaling of all moments
of the surface magnetization according to Eq. \ref{msavcriti}.
Our numerical data for the moments are compatible with Eq. \ref{msavcriti}
with an exponent
\begin{eqnarray}
x_s (d=3) \simeq 1.34
\label{xs3d}
\end{eqnarray}
that would correspond to
\begin{eqnarray}
a(d=3) = \frac{x_s}{\omega_c} -1 \simeq 4.5
\label{a3d}
\end{eqnarray}
and to the exponent (Eq. \ref{defbetas})
\begin{eqnarray}
\beta_s (d=3) = x_s \nu_{av} \simeq 1.76
\label{beta3d}
\end{eqnarray}
\section{ Conclusion }
\label{sec_conclusion}
Since the 'Cavity-Mean-Field' approximation developed for the Random Transverse Field Ising Model
on the Cayley tree \cite{ioffe,feigelman,dimitrova} has been found
to reproduce the known exact result for the surface magnetization in $d=1$ \cite{dimitrova},
we have proposed to extend these ideas in finite dimensions $d>1$
via a non-linear transfer approach for the surface magnetization.
In the disordered phase, the linearization (Eq \ref{milinear}) of the transfer equations
correspond to the transfer matrix for a Directed Polymer in a random medium
of transverse dimension $D=d-1$, in agreement with the leading order perturbative
scaling analysis \cite{transverseDP}.
We have presented numerical results of this non-linear transfer approach in dimensions $d=2$
and $d=3$, where large system sizes can be easily studied.
In both cases, we have found that the critical point is
governed by Infinite Disorder scaling. In particular exactly at criticality,
the one-point surface magnetization scales as $\ln m_L^{surf} \simeq - L^{\omega_c} v$,
where $\omega_c(d)$ coincides with the droplet exponent $\omega_{DP}(D=d-1)$
of the corresponding Directed Polymer model, with
$\omega_c(d=2)=1/3$ and $\omega_c(d=3) \simeq 0.24$. The distribution $P(v)$
of the positive random variable $v$ of order $O(1)$
presents a power-law singularity near the origin $P(v) \propto v^a$ with $a(d=2,3)>0$ so that all
moments of the surface magnetization are governed by the same power-law decay
$\overline{ (m_L^{surf})^k } \propto L^{- x_s}$ with $x_s=\omega_c (1+a)$
independently of the order $k$.
Our conclusion is thus that this non-linear transfer approach is able
to lead to Infinite Disorder scaling, that had been found previously
via Monte-Carlo in $d=2$ \cite{pich,rieger} and via Strong Disorder RG
in $d=2,3,4$ \cite{motrunich,fisherreview,lin,karevski,lin07,yu,kovacsstrip,
kovacs2d,kovacs3d,kovacsentropy,kovacsreview}.
Exactly at criticality, the presence of activated scaling $\ln m_L^{surf} \simeq - L^{\omega_c} v$
means that the linearization of Eq. \ref{milinear} is typically still valid
also at criticality (and not only in the disordered phase),
so that the identity $\omega_c=\omega_{DP}(D=d-1)$ can be understood.
The rare cases where this linearization is not valid at criticality is when
the positive random variable $v$ happens to be smaller than $1/L^{\omega_c}$.
Our conclusion is thus the following :
(i) in the disordered phase and for 'typical' situations exactly at criticality,
the linearization of Eq. \ref{milinear} is valid
and coincides with the leading order perturbative
scaling analysis \cite{transverseDP} : it is thus
likely to give exact values for critical exponents,
in particular for the exponent $\omega_c$ of activated scaling.
(ii) in the ordered phase and for 'rare' situations at criticality,
the non-linear terms of the transfer approach plays an important role.
Since they come from an uncontrolled approximation, the critical exponents
like $\beta_s$ and $x_s$ that are determined by these non-linear contributions
could be different from the exact ones.
To judge the accuracy of this approximation, it would be very helpful to compare
with other approaches like Quantum Monte-Carlo and Strong Disorder RG
(but up to now, these other approaches have not studied surface properties).
|
\section{Introduction}
\label{intro}
Investigating the physical nature and the dynamics of penumbral filaments is essential in order to understand the
structure and the evolution of sunspots and their surrounding moat regions. Many important observational
aspects of penumbral filaments are well settled down, although their interpretation is still a source of
debate, e.g., the brightness of penumbral filaments, the inward motion of bright penumbral grains, the
Evershed flow, the Net Circular Polarization (NCP), as well as moving magnetic features in the sunspot moat.
During the last few years, new observational discoveries, e.g., dark cored penumbral filaments
\citep{scharmeretal02}, strong downflow patches in the mid and outer penumbra \citep{ichimotoetal07a},
penumbral micro-jets \citep{katsukawaetal07, jrcakkatsu08}, or twisting motions of penumbral filaments
\citep{scharmeretal02,rimmelemarino06,ichimotoetal07b,ningetal09} broadened the number of unknowns and
gave new impulse to the investigation of sunspots. For recent reviews see \citet{borrero11}, \citet{bellot10}, \citet{borrero09},
\citet{sch09}, or \citet{tritschler09}. An especially controversial issue is the study of the twisting
motions of penumbral filaments. On the one hand, \citet{ichimotoetal07b} consider twisting motions as an
apparent phenomenon, produced by lateral motions of intensity fluctuations associated with overturning convection.
On the other hand, \citet{ryutovaetal08} propose that the observed twist is an intrinsic property of penumbral
filaments and is produced as a consequence of reconnection processes which take place in the penumbra.
\citet{suetal08,suetal10} conclude that the twist of penumbral filaments changes with time
caused by an unwinding process.
In the present paper we study the twist of filaments of the inner penumbra of a sunspot by means of the magnetic
helicity. We take advantage of the 3D geometrical model of a section of the inner penumbra of a sunspot described
in \citet{puschmannetal10a} [hereafter, Paper I]. We use observations of the active region AR 10953 near solar disk
center obtained on 1$^{\rm st}$ of May 2007 with the Hinode/SP. The inner, center side, penumbral area under study
was located at an heliocentric angle $\theta$\,=\,4.63$^{\circ}$. To derive the physical parameters of the solar
atmosphere as a function of continuum optical depth, the SIR (Stokes Inversion based on Response function)
inversion code \citep{ruizcobodeltoro92} was applied to the data set. The 3D geometrical model was derived by
means of a genetic algorithm that minimized the divergence of the magnetic field vector and the deviations from
static equilibrium considering pressure gradients, gravity and the Lorentz force.
We can not assess the unicity of the resulting model: the found solution just minimizes the divergency of
the magnetic field and the modulus of the net force, neglecting the contribution of the aceleration terms.
For a detailed description we refer to Paper I. In \citet{puschmannetal10b} [hereafter, Paper II], we calculated
the electrical current density vector $\vec{J}$ in the above mentioned area and found the horizontal component
of the electrical currents $\sim 4$ times larger than the vertical component ${J}_z$ \citep[thus confirming the
results of][]{pevtsovandperegud90, georgoulislabonte04}. In addition, we concluded that the magnetic field at the
borders of bright penumbral filaments departs from a force-free configuration \citep[see also][]{zhang10}.
These results are strongly significant considering that we have imposed that our solution minimizes the net force,
including the Lorentz force, and consequently, a force free solution should be found if it were possible.
\begin{figure}
\begin{center}
\includegraphics[scale=0.6]{Fig1.2.eps}
\end{center}
\caption{Left panel: Length $\ell$ of the field lines of $\vec{B}$ integrated from the top
($z=200$\,km) to the bottom layer ($z=0$\,km) or until $x=4.2$\,Mm.
Thick isolines of equal $\ell$ define 14 areas (flux tubes) related to spines (black) and intraspines (white),
respectively.
Thin isolines denote different thresholds for selecting the flux tube area.
Right panel: Vertical component of $\vec{B}$ at $z$\,=\,200\,km.
Contour lines correspond to horizontal cuts through the 14 flux tubes with the largest area
at $z=200, 150, 100, 50, 0$\,km (the thickness of each line diminishes with depth). Black and white contours
distinguish between flux tubes related to spines and intraspines, respectively.}
\label{Fig1}
\end{figure}
We can evaluate the magnetic field lines by the integration of $\vec{B}$ starting at each pixel
of the top layer of our volume of the inner penumbra (of 4.2 Mm $\times$ 5.6 Mm $\times$
0.2 Mm, see Paper II). In the left panel of Fig.~\ref{Fig1} we present the length $\ell$
of each field line. As our sunspot has negative polarity, $\vec{B}$ points downwards and
we thus integrate the field lines from the top layer to the bottom layer. The majority of field
lines end up at the bottom layer, except the lines starting at larger X coordinates
in our FOV.
In the whole analized volume the magnetic field has the same polarity, and consequently,
the field lines do not present maxima nor minima in our region, i.e., the field lines always travel
downwards. This fact, as we will see later, simplifies the evaluation of the magnetic helicity.
Areas with larger $\ell$ correspond to intraspines, since $\vec{B}$ is more horizontal, the field
lines thus traverse larger distances inside our volume.
In the left panel of Fig.~\ref{Fig1}, we selected 14 areas, 7 corresponding to intraspines (thick white
contours) and 7 to spines (thick black contours), according to the length $\ell$ of the magnetic field lines.
For each of the selected zones we define a volume delimited by the field lines setting off from each
pixel of the closed curve of the top layer.
In the right panel of Fig.~\ref{Fig1} we show the vertical component of $\vec B$ at
$z$\,=\,200\,km together with horizontal cuts through each of these volumes at
$z=200, 150, 100, 50, \&\,0$\,km. Since the umbra is placed at the right hand side in
the FOV, the cuts at deeper layers are displaced to the right. Finally we checked that the
magnetic flux traversing each of the cuts is approximately constant:
the standard deviation of the relative variation of the magnetic flux between the top
and the bottom layer is 4.5\%. Thus, each volume can be approximately
considered as a flux tube.
However, the 14 areas were selected quit arbitrarily. In order to study the dependence
of twist, writhe and magnetic helicity on the flux tube area, 40 additionally smaller
tubes have been defined inside the larger
ones (thin isolines in the left panel of Fig.~\ref{Fig1}). Thus, for most
of the 14 zones we have several tubes of different size, many of
them being the internal part of the larger
one. Consequently, we have 54 tubes, 27 of them
related to intraspines and 27 to spines.
\section{Magnetic helicity, Twist, and Writhe}
The study of helicity of solar magnetic features has been a hot topic during at least the last 25 years.
Magnetic helicity has been investigated in solar structures at different spatial scales in the
photosphere and chromosphere, as well as in the solar wind \citep[see e.g. the reviews of]
[and references therein]{brownetal99, rust02, pevtsovandbala03, demoulin07, demoulinandpariat09}.
The helicity in penumbral filaments has been analyzed by means of some proxies by, e.g.,
\citet{ryutovaetal08}, \citet{tiwariandvenkata09}, \citet{suetal10}, and \citet{zhang10}.
\begin{figure*}
\begin{center}
\includegraphics[scale=1.3]{Fig2.3.eps}
\end{center}
\caption{From left to right: local twist ($T^{\mbox{\rm\tiny{loc}}}$), $\alpha_{z}/4\pi$,
current helicity density ($h_{c}$), and $4h_{c_{z}}$ evaluated at $z=200$\,km.
The isolines are the same as in the left panel of Fig.~\ref{Fig1}.}
\label{Fig2}
\end{figure*}
The magnetic helicity, $H_{m}$, quantifies how the magnetic field is twisted,
writhed, and linked.
$H_{m}$ plays a key role in magneto-hydrodynamics because
it is almost conserved in a plasma having a high magnetic Reynolds number \citep[see e.g.,][]
{berger84}. The magnetic helicity of a vector field $\vec{B}$, fully contained within a volume
$\mathcal{V}$ and bounded by a surface $\mathcal{S}$ (i.e., the normal component
$B_{n}=\vec{B}\cdot\vec{n}$ vanishes at any point of $\mathcal{S}$), is \citep{elsasser56}:
\begin{equation}
H_{m}=\int_{\mathcal{V}}\vec{A}\cdot\vec{B}\,\, \mathrm{d}^{3}x,
\label{eq1}
\end{equation}
where the vector potential $\vec{A}$ satisfies $\vec{B}=\nabla \times \vec{A}$.
\citet{bergerandfield84} showed that Eq.~\ref{eq1},
is not gauge-invariant if the volume of interest is not bounded by a magnetic surface, i.e., if $\vec{B}$
crosses $\mathcal{S}$ (as in the case of the volume of the penumbra retrieved from our observations).
In this case the relative magnetic helicity \citep{finnandandantonsen85} should be used:
\begin{equation}
H_{m}^{rel}=\int_{\mathcal{V}}(\vec{A}+\vec{A}_{p})\cdot(\vec{B}-\vec{B}_{p})\,\, \mathrm{d}^{3}x,
\label{relativehm}
\end{equation}
where $\vec{B}_{p}$ is a potential field having the same normal component $B_{n}$ on $\mathcal{S}$,
and $\vec{A}_{p}$ is its vector potential.
The relative helicity reflects twist, writhe, and linkage with respect to a current-free (potential) field,
i.e., its minimum-energy state for the given $B_{n}$-condition on $\mathcal{S}$.
The relative magnetic helicity so defined is gauge-invariant
and has the same conservation properties and amount of topological information
as the magnetic helicity.
Throughout this article, the term magnetic helicity refers to the relative magnetic helicity. For an isolated
magnetic flux rope, $H_{m}^{rel}$ is proportional to the sum of its twist $T$ and writhe $W$
\citep[][]{bergerandfield84, Toroketal10}:
\begin{equation}
H_{m}^{rel}=(T+W)\Phi^{2}
\label{TandW}
\end{equation}
where $\Phi$ is the magnetic flux of the rope. The twist is the turning angle of a bundle of
magnetic field lines around its central axis, whereas the writhe quantifies the helical deformation
of the axis itself. Following \citet{bergerandprior06}, the twist of an infinitesimal rope is given by
$T=\int T^{\mbox{\rm\tiny{loc}}}{\mathrm{d}l}$, $l$ being the arc length along the central
field line of the rope and $T^{\mbox{\rm\tiny{loc}}}$ the local twist:
\begin{equation}
T^{\mbox{\rm\tiny{loc}}}=\frac{\mathrm{d}T}{\mathrm{d}l}=\frac{\mu_{0}J_{||}}{4 \pi B_{||}}\,\,,
\label{local}
\end{equation}
being $J_{||}$ and $B_{||}$ the components of the current and magnetic field parallel to the central
field line of the rope. With this definition, $T=1$ when the field lines twist around the axis by an
angle of $2\pi$ and $T^{\mbox{\rm\tiny{loc}}}$ is the local twist per unit length evaluated at a given
geometrical height at each pixel.
If the rope has a non-infinitesimal cross section $\Sigma$, the local twist of the rope
$T^{\mbox{\rm\tiny{loc}}}_i$ is given by the average of the infinitesimal local twist,
$T^{\mbox{\rm\tiny{loc}}}$ over $\Sigma$.
\citet[][see also \citet{Toroketal10}]{bergerandprior06} give expressions for the writhe of specific
geometrical configurations.
Provided that the magnetic field lines in the inner penumbral region studied in this paper always
travel downwards in $z$, i.e. without showing maxima nor minima between the $z_0=$200\,km and $z_1=$0\,km
height layers, we can evaluate the writhe using a very simplified formula:
\begin{equation}
W=\frac{1}{2 \pi}\int_{z_0}^{z_1}\frac{1}{1+|\tau_z|}{(\vec{\tau} \times \vec{\tau'})}_{z}{\mathrm{d}z}\,,
\label{writhe}
\end{equation}
obtained as a particular case of the more general formula of \citet{bergerandprior06}.
In Eq.~\ref{writhe}, $\vec{\tau}$ stands for the tangent vector to the tube axis;
${\tau_z}$ for the vertical component of $\vec{\tau}$; and
$\vec{\tau'}=\frac{\mathrm{d}\vec{\tau}}{\mathrm{d}z}$.
Equations~\ref{TandW}, \ref{local}, \& \ref{writhe} allow the evaluation of the twist,
the writhe and the magnetic helicity of an isolated tube.
What happens in the case of a non-isolated tube, as is clearly the case of the penumbral tubes?
The writhe of a magnetic rope, isolated or not, is a measure of the helical deformation of
the axis of the rope, while the twist quantifies the winding of the magnetic field lines
of the rope around its axis. The magnetic helicity measures the linking number of the
field lines, averaged over all pairs of lines, and weighted by the flux \citep{bergerandprior06, Moffatt69}.
We can simplify the case of a non-isolated tube to a scenario
in which we have just two adjoining tubes. It is clear that we can define the writhe and
twist for each individual tube, and consequently its magnetic helictiy, but the helicity of the
whole configuration is not just the sum of both contributions. We would need to include an extra
term taking into account the linking between both tubes: the mutual helicity.
Consequently the application of equations~\ref{TandW}, \ref{local}, \& \ref{writhe}
to our tubes retrieves only the contribution of the local values of the twist, writhe and the
self helicity. The contribution of the surrounding tubes to the helicity
of each flux tube is not considered. This contribution, the mutual helicity, could
be larger than the self helicity \citep[see e.g.,][]{renierpriest07}.
The mutual helicity can be calculated using the procedure described
in \citet{bergerandprior06}. However, for the scope of this paper we limit the calculation
to self helicity.
\section{Proxies of the magnetic helicity}
The current helicity density is defined \citep[see e.g.,][]{seehafer90} as
$h_{c}={\vec B}\cdot\nabla\times{\vec B}$.
If we take as magnetic ropes the tubes defined by the field lines starting in each pixel,
(being then $B_{||}=B$), Eq.~\ref{local} becomes $T^{\mbox{\rm\tiny{loc}}} = h_{c} / 4 \pi B^{2}$.
The parameter $\alpha$ is usually defined in a force-free configuration, i.e., when ${\vec B}$ is
parallel to its curl, by $\nabla\times{\vec B}$\,=\,$\alpha {\vec B}$.
The $\alpha$ parameter can be defined
for non force-free fields: $\alpha= {\vec B}\cdot\nabla\times{\vec B}/B^{2}$. This is the
definition\begin{footnote}{\citet{yeatesetal08} denominate current helicity to this generalized
$\alpha$ parameter.}\end{footnote} we will use throughout the paper.
In order to see how this re-defined
$\alpha$ differs from the force-free definition, let us decompose
$\nabla\times{\vec B}=(\nabla\times{\vec B})_{||} + (\nabla\times{\vec B})_{\perp}$.
From its definition,
$\alpha$ becomes equal to
$\alpha=\pm|(\nabla\times{\vec B})_{||}|/B$ which is equal to the classical
definition for a force-free field. The $\pm$ is needed to consider the case
when $(\nabla\times{\vec B})_{||}$ and $\vec B$ point in opposite direction, i.e.,
when $\alpha$ is negative.
For a no force free-field, the parameter $\alpha$ is then
the ratio between the parallel component of the curl of the magnetic field and its modulus.
Evidently, in a general case, we will
have $\alpha=h_{c}/B^{2}=4\pi\,T^{\mbox{\rm\tiny{loc}}}$.
Given the difficulty of empirically obtaining $H_{m}^{rel}$ and ${\vec J}$, one finds many works where
different proxies were used. Before evaluating the magnetic helicity we can calculate, from
our data, some of the most usual proxies of the magnetic helicity. Among them the most common
proxies, with several different but more or less equivalent definitions, are the
$\alpha_{z}=(\nabla\times{\vec B})_{z}/B_{z}$ parameter
\citep[see e.g.,][and references therein]{suetal09, suetal10, pevtsovetal08}, and the
parameter $h_{c_{z}}=B_{z}(\nabla\times{\vec B})_{z}$ \citep[see e.g.,][]{zhang10}. It is evident
that for a force-free field $\alpha_{z}=\alpha$ and $h_{c_{z}}=\alpha B_{z}^2=h_{c}B_{z}^2/B^2$.
That means that
the $h_{c_{z}}$ parameter could be meaningless for nearly horizontal magnetic fields,
such as those found in sunspot penumbrae.
Many authors suppose that
the sign of the integral of $\alpha_{z}$ (or $h_{c_{z}}$) over the volume of a magnetic structure
coincides with the sign of $H_{m}^{rel}$, although this fact has not yet been demonstrated
\citep{demoulin07}. On the other hand, \citet{hagyardpevtsov99} point out that $h_{c_{z}}$ only
considers the vertical component $J_{z}$ of the
electric current density vector, $h_{c_{z}}$ can
strongly differ from $h_{c}$, provided that $J_{z}$ is much smaller than the horizontal
components, at least in the inner penumbra (see Paper II). Besides, \citet{pariatetal05} comment
that, since the magnetic helicity is a global quantity, it is not obvious that a helicity density
has any physical meaning.
We evaluated these proxies in the inner penumbral region under study.
In the left panels of Fig.~\ref{Fig2} we present $T^{\mbox{\rm\tiny{loc}}}$ (evaluated from Eq.~\ref{local})
and $\alpha_{z}/4\pi$ evaluated at each pixel at $z=200$\,km. As in the
inner penumbra the magnetic field is not force-free at the borders of bright penumbral filaments
(see Paper II), $\alpha_{z}/4\pi$ (2$^{\rm nd}$ panel) is only qualitatively similar to
$T^{\mbox{\rm\tiny{loc}}}$. In the 3$^{\rm rd}$ and 4$^{\rm th}$ panel we present $h_{c}={\vec B}\cdot\nabla\times{\vec B}$
and $h_{c_{z}}$ evaluated at $z=200$\,km. $h_{c_{z}}$ is multiplied by a factor 4 just to make
easier its comparison with $h_{c}$. As we have seen before,
$T^{\mbox{\rm\tiny{loc}}} = h_{c} / 4 \pi B^{2}$, and thus the general aspect of $h_{c}$ is
very similar to $T^{\mbox{\rm\tiny{loc}}}$. However, $h_{c_{z}}$
resembles $h_{c}$ only marginally (the standard deviations are
$\sigma(h_{c})=2.1\,G^{2}\,m^{-1}$ and $\sigma(h_{c_{z}})=0.4\,G^{2}\,m^{-1}$).
The values of $\alpha_{z}/4\pi$ and $h_{c_{z}}$ at $z=200$\,km obtained here are
very similar to the results found in the literature: \citet{tiwariandvenkata09} found
that $\alpha_{z}/4\pi$ varies around $\pm 0.15 Mm^{-1}$ along azimuthal paths in the
middle penumbra; \citet{suetal10} found a fluctuation of $\alpha_{z}/4\pi$ larger than
$\pm 0.05 Mm^{-1}$ over an inner penumbral region; \citet{suetal09} found that $h_{c_{z}}$
fluctuates along an azimuthal path in the inner penumbra with an amplitude larger than
1\,$G^{2}\,m^{-1}$ while \citet{horstpeter08} found penumbral mean values of about
0.04\,$G^{2}\,m^{-1}$. In the four panels of Fig.~\ref{Fig2}, the outlined areas were
selected by different thresholds of $\ell$, the length of the magnetic field lines
between the layers $z=200$\,km and $z=0$\,km (see also Section~\ref{intro} and Fig.~\ref{Fig1}).
The sign of the integral of $T^{\mbox{\rm\tiny{loc}}}$ and $\alpha_{z}/4\pi$ over
the above mentioned areas only coincides in 39\% of the 54 tubes (if we consider only the areas related
to the intraspines this value decreases to 15\% of the 27 tubes). The same figures are obtained for the
percentage of coincidence between the signs of the integrals of $h_{c}$ and $h_{c_{z}}$, approximately.
This weak coincidence demonstrates that, at least in the inner penumbra of a sunspot,
$\alpha_{z}/4\pi$ and $h_{c_{z}}$ are not good estimates of
$T^{\mbox{\rm\tiny{loc}}}$ and $h_{c}$, respectively. As already shown in Paper II, the magnetic field in
the area under study significantly departures from a force-free configuration.
Note that $T^{\mbox{\rm\tiny{loc}}}$
reaches significant values only at the borders of the intraspines: these are exactly the
areas where the electric current density is large (see Fig. 1 of Paper II).
Furthermore, often $T^{\mbox{\rm\tiny{loc}}}$ changes its sign at both sides of bright
filaments, i.e., for the majority of the intraspines, $T^{\mbox{\rm\tiny{loc}}}$ shows
negative values at the upper (larger Y-coordinate) borders of the filaments and positive
values at the lower borders. The alternation of the sign in the twist is also observed
(although less evident) in the maps of $\alpha_{z}$ and $h_{c_{z}}$ and it is clearly
visible in \citet{tiwariandvenkata09}, \citet{suetal09}, \citet{suetal10}, and \citet{zhang10}.
This phenomenon can be explained if we consider that the field lines of the magnetic background
component wrap around the intraspines \citep{borreroetal08} and tend to meet above the
intraspines, thus generating a curvature of different sign in the field lines at both
sides of the intraspines. Thus $T^{\mbox{\rm\tiny{loc}}}$ could be measuring the twist of
the background field wrapping around the intraspines rather than the twist of the field lines
of the intraspines themselves. However, the alternation of signs of the twist at both sides
of a penumbral filament is compatible with the magnetohydrostatic equilibrium model of a
magnetic flux tube built by \citet{borrero07}. This model includes a transverse component
of $\vec{B}$ having opposite twist at both sides of a plane longitudinally cutting the flux
tube. This model is able of explaining both the dark cored penumbral filaments and the net
circular polarization observed in penumbral filaments \citep{borreroetal07}. \citet{magara10}
suggests the existence of an intermediate region where the magnetic field has a
transitional configuration between a penumbral flux tube and the background field: in such areas,
coinciding with the largest electrical current density (see Paper II), penumbral micro-jets
are produced as observed by \citet{katsukawaetal07}.
\section{Numerical test}
To check its correctness, the procedure used for the evaluation of twist, writhe, and magnetic
helicity was applied to two different analytical cases.
In the first case we consider that the magnetic field lines follow a helix around a vertical
straight line. The magnetic field vector is defined by
$\vec{B}=B_0\,\hat{z} + B_1\,r\,\hat{\theta} $
with $B_0$ and $B_1$ being constant.
$\vec{B}$ could easily be decomposed in a potential $\vec{B}_p=B_0\,\hat{z}$
and a close (toroidal) field $\vec{B}_c=B_1\,r\,\hat{\theta}$.
Obviously, the potential component $\vec{B}_p$ fulfills the conditions required
in Eq.~\ref{relativehm} in the case of a cylindrical tube: $\vec{B}$ and $\vec{B}_{p}$
have the same normal component on the external surface of the tube.
The determination of the vector potential, in this case, is straightforward:
$\vec{A}=B_0\,r/2\,\hat{\theta} - B_1\,r^2/2\,\hat{z}$.
The vector potential of the potential component will be
$\vec{A_p}=B_0\,r/2\,\hat{\theta}$.
Using Eq.~\ref{relativehm}, the magnetic helicity of a cylindrical
tube of height $L$ and radius $R$ becomes
$H_m=\frac{1}{2}\pi\,B_0\,B_1\,R^4\,L$.
The magnetic flux of this tube is $\Phi=\pi\,B_0\,R^2$.
This tube has a zero writhe because its axis is a straight line.
From Eq.~\ref{TandW}, the twist follows as
\begin{equation}
T=\frac{B_1\,L}{2\pi\,B_0}.
\label{Analitic_Twist}
\end{equation}
This is obviously the expected result, provided that the pitch (of screw-step) of our helix is
$\frac{2\pi\,B_0}{B_1}$ and the twist is a measure of the number of turns done by
the magnetic field lines along a longitude $L$.
In the first four rows of Table \ref{table1} we present the analytical
(i.e., using Eq.~\ref{Analitic_Twist}) and numerical results (using Eqs.~\ref{local}
and \ref{writhe}) for tubes with $B_0=-0.2\, {\mbox{T}}$, $B_1=0.04\,
{\mbox{T}}\,{\mbox{Mm}}^{-1}$, $L=0.225\,{\mbox{Mm}}$ and a radius $R$ equal to $0.2$
and $1$ Mm. We used the same spatial grid as in the observational case.
In order to check the accuracy of the determination of twist, writhe and magnetic helicity in
a more general case, we carried out a second test, building a helical magnetic field that
turns around a helical axis. Let us suppose that the axis of the tube is a vertical
helix of radius $R_h$ turning an angle $\Psi$ through a length $L=0.22\,{\mbox{Mm}}$. Then, the
magnetic field at the axis will be $\vec{B}=B_0\, \hat{z} + B_1\,R_h\, \hat{\theta} $ with
$B_1=B_0\,\Psi/L$. Following \citet{bergerandprior06}, the writhe of a magnetic flux tube
whose axis is a helix can be easily evaluated in terms of its polar writhe (i.e., area/$2\pi$ of
the section of the unity sphere limited by the tantrix curve and the north pole.
The tantrix curve is the path, the tip of the tangent vector takes on the unit sphere).
In our case, the writhe becomes:
\begin{equation}
W=\frac{\Psi}{2\pi}(1-\frac{B_0}{|B|}).
\label{Analitic_Writhe}
\end{equation}
Once we have the axis, we can easily build a tube with a given twist
around such an axis. We chose the radius of the tube as $R=0.2$ Mm, and the radius of
the helical axis as $R_h=0.4\,{\mbox{Mm}}$, and $B_1=0.5\,{\mbox{T\,Mm$^{-1}$}}$. The angle $\Psi$
takes a value of $9.124$ degrees in order to have an analytical writhe
(i.e., using Eq.~\ref{Analitic_Writhe}) of $0.001$.
The added twist takes the values $0.0$, $0.001$, $0.01$, $0.1$, and $-0.001$.
Note that, as the magnetic helicity is proportional to the sum of twist and writhe, in
the last case we will have a null magnetic helicity. The results of these tests are
presented in the ten bottom rows of Table \ref{table1}.
\begin{table}
\caption{Test results: Writhe ($W$), twist ($T$), and magnetic helicity ($H_m$), for a helical
magnetic field (first four rows) and a helical magnetic field winding around a helical axis.}
\label{table1}
\begin{center}
\leavevmode
\begin{tabular}{ccccc} \hline \hline
$R$ [Mm] & & $W$ & $T$ & $H_m\,\,[{\mbox {Mx}}^2]$ \\ \hline
0.2 & analytical& 0.00 & -7.16e-3 & -4.52e+34 \\
& numerical & -9.e-28 & -7.16e-3 & -4.50e+34 \\ \hline
1.0 & analytical& 0.00 & -7.16e-3 & -2.83e+37 \\
& numerical & -9.e-28 & -7.16e-3 & -2.81e+37 \\ \hline
0.2 & analytical& 1.00e-3 & 0.00 & 8.16e+34 \\
& numerical & 0.99e-3 & 1.90e-6 & 8.07e+34 \\ \hline
0.2 & analytical& 1.00e-3 & 1.00e-3 & 1.63e+35 \\
& numerical & 0.99e-3 & 1.07e-3 & 1.67e+35 \\ \hline
0.2 & analytical& 1.00e-3 & 1.00e-2 & 8.97e+35 \\
& numerical & 0.98e-3 & 1.06e-2 & 9.47e+35 \\ \hline
0.2 & analytical& 1.00e-3 & 1.00e-1 & 7.01e+36 \\
& numerical & 0.97e-3 & 0.99e-1 & 6.91e+36 \\ \hline
0.2 & analytical& 1.00e-3 & -1.00e-3 & 0.00 \\
& numerical & 0.99e-3 & -1.06e-3 & -6.14e+33 \\ \hline
\end{tabular}
\end{center}
\end{table}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5]{Fig3.1.eps}
\end{center}
\caption{Panel (a): Length of the axis of each flux tube. The values
corresponding to tubes of the same zone are connected by lines. Cross symbols correspond to intraspines and small circles to spines.
Panels (b), (c), and (d): writhe, twist, sum of writhe and twist.}
\label{Fig3}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[scale=.50]{Fig4.2.eps}
\end{center}
\caption{Panels (a) and (b): normalized twist ($T/\zeta$) and writhe ($W/\zeta$) {\it versus}
the length of the axis of each flux tube $\zeta$. The absolute value $|W/\zeta|$ has been
overplotted in panel (b) with triangle symbols. The dashed line is the linear fit of
$|W/\zeta|$.
Panel (c): wavenumber.
Panel (d): $T/\zeta$ versus $\alpha/4\pi$ averaged over the section of each flux tube at $z=200$\,km.
Panels (e) and (f): $T$ versus the average of $\alpha_z/4\pi$ and $h_{c_{z}}$
respectively. In panels (d), (e) and (f) the straight line with slope 1 has been overplotted.
Cross symbols correspond to intraspines and small circles to spines.
}
\label{Fig4}
\end{figure}
\section{Results}
\begin{table}
\caption{Axis length $\zeta$, magnetic flux $\Phi$, twist, writhe and
magnetic helicity for the 14 largest zones. The 7 first (last) rows are related to
intraspines (spines).}
\label{table2}
\begin{center}
\leavevmode
\begin{tabular}{cccccc} \hline \hline
index& $\zeta$ [Mm] & $\Phi$ [Mx] & T & W & $H_m\,\,[{\mbox {Mx}}^2]$ \\
~ 0 & 0.79 & -3.05e+18 & -0.0138 & ~ 0.0258 & ~ 1.11e+35\\
~ 4 & 0.80 & -1.56e+18 & -0.0272 & ~ 0.0182 & -2.19e+34\\
~ 9 & 1.06 & -4.71e+18 & -0.0210 & ~ 0.0024 & -4.12e+35\\
15 & 0.92 & -2.55e+18 & -0.0224 & -0.0163 & -2.51e+35\\
20 & 0.80 & -5.69e+18 & ~ 0.0111 & ~ 0.0041 & ~ 4.94e+35\\
23 & 1.02 & -2.64e+17 & -0.0054 & ~ 0.0166 & ~ 7.80e+32\\
25 & 1.00 & -5.13e+17 & ~ 0.0096 & ~ 0.0079 & ~ 4.59e+33\\
\hline
28 & 0.32 & -4.43e+18 & -0.0023 & ~ 0.0000 & -4.42e+34\\
29 & 0.33 & -3.41e+18 & -0.0028 & ~ 0.0021 & -8.34e+33\\
30 & 0.34 & -6.29e+18 & -0.0041 & ~ 0.0011 & -1.20e+35\\
36 & 0.34 & -7.73e+18 & -0.0065 & ~ 0.0010 & -3.30e+35\\
40 & 0.37 & -6.30e+18 & -0.0041 & ~ 0.0015 & -1.01e+35\\
41 & 0.36 & -1.13e+19 & -0.0001 & ~ 0.0007 & ~ 7.56e+34\\
46 & 0.35 & -2.04e+19 & -0.0010 & -0.0017 & -1.15e+36\\
\hline
\end{tabular}
\end{center}
\end{table}
In Table~\ref{table2} we present the resulting values of the axis length, magnetic flux,
twist, writhe and self magnetic helicity for the flux tubes of the 14 largest zones. As the
magnetic helicity depends on the square of the magnetic flux, and our selected areas are
very different in area, the resulting magnetic helicity varies over a wide range of several
orders of magnitude. To study the dependence of the precedent quantities on the length of
the respective axis, in
Fig.~\ref{Fig3}, we plot the length of the axis of each flux tube $\zeta$, the writhe,
the twist, and the sum of twist and writhe for the $54$ selected tubes related to intraspines
(index ranging form $0$ to $26$) and spines (index ranging form $27$ to $53$). The values corresponding
to tubes of the same zone (see left panel of Fig.~\ref{Fig1}) are connected by straight lines.
As the magnetic field in the spines is more vertical than in the intraspines, the length of the axis
of the tubes related to the
spines is clearly shorter. For each intraspine/spine zone the length of the axis grows with the index
because each of the related tubes was chosen in the interior of the preceding one. The
writhe of the intraspines does not follow a clear pattern, but most of the tubes have a positive
writhe. The spines show a nearly zero writhe. The twist of the intraspines, however, takes
nearly always negative values. Consequently,
the twist is partially canceled by the writhe, thus the absolute value of the self magnetic helicity is,
nearly always, lower than the absolute value of the twist.
In panels (a), and (b) of Fig.~\ref{Fig4} we plot the twist, and writhe per unit
length versus the length of the axis of each flux tube $\zeta$. The normalized twist does
not show a clear dependence with the length of the axis, but the absolute value of the
normalized writhe in the intraspines clearly decreases with increasing $\zeta$. The axes of
tubes related to intraspines are less wrung when the tubes are more horizontal
(i.e., in the central part of the intraspines). As the writhe of the
spines is very small, we can conclude that the writhe reaches only significant values
when the tube includes the border of a intraspine.
In panel (c) we plot the wavenumber (1/$pitch$), i.e., the number of turns done by the magnetic
field per length unit as a function of $\zeta$. Flux tubes related to spines show slightly smaller
values but any clear dependence is not observed.
In panel (d) we plot the normalized twist versus $\alpha/4\pi$ (i.e., the local twist
$T^{\mbox{\rm\tiny{loc}}}$ at each pixel) evaluated at $z=200\,$km and averaged over
each structure.
Given the good correlation between both magnitudes we can use the average of the local
twist as a good proxy for the normalized twist. Consequently, we can explain the small
obtained twist values in terms of the local twist: as we have seen in Fig.~\ref{Fig2},
$T^{\mbox{\rm\tiny{loc}}}$ reveals significant values with opposite sign at the borders
of the penumbral filaments, leading to a cancellation of the twist when integrating over
the filament.
In order to assess the reliability of the most used proxies we
plot, in panels (e) and (f), the average twist of each structure versus the average
value of $\alpha_z/4\pi$ and $h_{c_{z}}$, respectively. Both proxies are very bad indicators
of the average twist of intraspines but they give a qualitatively good
agreement (better in the case of $h_{c_{z}}$) for the spines. This asymmetry
could be explained by the fact that both $\alpha_z$ and $h_{c_{z}}$
are only related with
the vertical component $J_{z}$ of the electric current density vector,
which, as we show in Paper II, is much smaller than the horizontal
components, mainly in and around the intraspines. On the other hand,
in a force-free configuration $\alpha_z=\alpha$ and $h_{c_{z}}=h_c$,
and then, the discrepancy between these parameters is a clear result of
the non-validity of the force-free approximation in the inner penumbra:
In Paper II we have already shown that, at the borders of bright penumbral
filaments, the magnetic field strongly departures from a force-free configuration.
\section{Conclusions}
In the present work we calculated the parameter $\alpha$ and its proxy $\alpha_z$, the
current helicity density $h_{c}$ and its proxy $h_{c_{z}}$, the twist, the writhe, and
the magnetic helicity of different structures of the inner penumbra of a sunspot.
The parameters are evaluated from a three-dimensional geometrical model obtained after
the application of a genetic algorithm on inversions of spectropolarimetric data observed
with {\it Hinode} \citep[see][Paper I]{puschmannetal10a}. We demonstrate, that in the inner penumbra the frequently
used proxies $\alpha_z$ and $h_{c_{z}}$ are only qualitative indicators of the local twist
(twist per unit length, evaluated under the assumption that the axis of a flux tube is
parallel to the magnetic field) of penumbral structures. As shown in \citep[][Paper II]{puschmannetal10b}, the
magnetic field in the area under study many times departs significantly from a force-free configuration
and the horizontal component of the electrical current density is significantly larger than the vertical one.
The local twist shows only significant values at the borders of bright penumbral filaments and reveals
opposite sign at each side of the bright filaments. The opposite sign might be the reason
for a cancellation of the twist when integrating over the filament, thus the twist of
the penumbral structures is very small. Significant values of the local twist are exactly
related to areas where the electric current density is large. The local twist could be
measuring the twist of the background field wrapping around the intraspines and/or the
twist of the field lines of the intraspines themselves; in the latter case the internal structure
of the tube would consist in two "cotyledons" (at both sides of a vertical plane
longitudinally cutting the tube), harboring each one a magnetic field of opposite
twist, compatible with the MHS model of \citet{borrero07}.
The writhe per unit length diminishes with increasing length (decreasing inclination) of the
axis of flux tubes related the intraspines. The small amount of twist and writhe shown
by the spines indicates that the background field lines, in these zones, are nearly straight.
A future study should clarify if the helicity apparent in the intensity maps of penumbral
filaments in the mid and outer penumbra of sunspots is produced by helical flux tubes
with a strong writhe or just by spurious effects produced by lateral intensity fluctuations.
In any case, it is clear that a twisted tube does not {\it per se} generate any intensity
fluctuation similar to the observations by \citet{ryutovaetal08}. Rather, there is the
necessity of a writhe of the tube, in such a way that different longitudinal portions
of the tube were at different optical depths producing changes in the observed intensity.
We will extend the present work (and necessarily the work presented in Paper I and II)
on the entire sunspot.
\acknowledgments
We thank the referee for fruitful comments.
Hinode is a Japanese mission developed and launched by ISAS/JAXA, collaborating with
NAOJ as a domestic partner, NASA and STFC (UK) as international partners. Scientific
operation of the Hinode mission is conducted by the Hinode science team organized at
ISAS/JAXA. This team mainly consists of scientists from institutes in the partner
countries. Support for the post-launch operation is provided by JAXA and NAOJ (Japan),
STFC (U.K.), NASA, ESA, and NSC (Norway).
Financial support by the Spanish Ministry of Science and Innovation
through projects AYA2010--18029, ESP 2006-13030-C06-01, AYA2007-65602, and the European
Commission through the SOLAIRE Network (MTRN-CT-2006-035484) is gratefully acknowledged.
The National Solar Observatory (NSO) is operated by the Association of Universities
for research in Astronomy (AURA), Inc., under a cooperative agreement with the National
Science Foundation. We thank V. Mart\'inez Pillet, C. Beck, and H. Balthasar for fruitful
discussions.
|
\section{Introduction}
\label{i}
The concept of \emph{weak solution} in fluid dynamics was
introduced in the seminal paper by Leray \cite{LER} in the context
of incompressible, linearly viscous fluids. Despite a concerted
effort of generations of mathematicians, the weak solutions still
represent the only available framework for studying the time
evolution of fluid systems subject to large data on long time
intervals, see Fefferman \cite{Feff}. The original ideas of Leray
have been put into the elegant framework of generalized
derivatives (distributions) and the associated abstract function
spaces of Sobolev type, see Ladyzhenskaya \cite{LAD}, Temam
\cite{TEM}, among many others. More recently, Lions \cite{LI4}
extended the theory to the class of barotropic flows (see also
\cite{EF70}), and, finally, the existence of global-in-time
generalized solutions for the full Navier-Stokes-Fourier system
was proved in \cite{EF70}, \cite[Chapter III]{FENO6} (see also
Bresch and Desjardins \cite{BRDE}, \cite{BRDE1} for an altrenative
approach based on a specific relation satisfied by the
density-dependent viscosity coefficients). Despite their apparent
success in the theory of \emph{existence}, the weak solutions in
most of the physically relevant cases are not (known to be)
uniquely determined by the data and may exhibit other rather
pathological properties, see Hoff and Serre \cite{HOSE}.
A fundamental test of \emph{admissibility} of a class of weak
solutions to a given evolutionary problem is the property of
\emph{weak-strong uniqueness}. More specifically, the weak
solution must coincide with a (hypothetical) strong solution
emanating from the same initial data as long as the latter exists.
In other words, the strong solutions are unique within the class
of weak solutions. {This problem has been intensively studied for the \emph{incompressible} Navier-Stokes system
for which, loosely speaking and simplifying, this is the ``best''
property that one can prove in this respect for the weak
solutions, see Prodi \cite{PR}, Serrin \cite{Serrin} for the first results in
this direction, and Escauriaza, Seregin, \v Sver\'ak \cite{ESSV}
for the most recent development.}
The main goal of the present paper is
to show that the class of weak solutions to the \emph{full}
Navier-Stokes-Fourier system introduced in \cite{FENO6} enjoys
the weak-strong uniqueness property.
The motion of a general compressible, viscous, and heat conducting fluid is described by means of three basic state variables: the mass density $\vr=\vr(t,x)$, the velocity field
$\vu = \vu(t,x)$, and the absolute temperature $\vt = \vt(t,x)$, where $t$ is the time, and $x \in \Omega \subset R^3$ is the space variable in the Eulerian coordinate system. The time evolution of these quantities is governed by a system
of partial differential equations - mathematical formulation of the physical principles of balance of mass, momentum, and entropy:
\bFormula{i1}
\partial_t \vr + \Div (\vr \vu) = 0,
\eF
\bFormula{i2}
\partial_t (\vr \vu) + \Div (\vr \vu \otimes \vu) + \Grad p(\vr, \vt) = \Div \tn{S}(\vt, \Grad \vu),
\eF
\bFormula{i3}
\partial_t (\vr s(\vr, \vt)) + \Div (\vr s(\vr, \vt) \vu) + \Div \left( \frac{ \vc{q}(\vt, \Grad \vt)}{\vt} \right) = \sigma,
\eF
where $p$ is the pressure, $s$ is the (specific) entropy, and where we have deliberately ommited the influence of external sources. Furthermore, we suppose that the viscous stress $\tn{S}$ is a linear function of the velocity gradient therefore described by Newton's law
\bFormula{i4}
\tn{S}(\vt, \Grad \vu) = \mu(\vt) \Big( \Grad \vu + \Grad^t \vu - \frac{2}{3} \Div \vu \tn{I} \Big) + \eta (\vt) \Div \vu \tn{I},
\eF
while
$\vc{q}$ is the heat flux satisfying Fourier's law
\bFormula{i5}
\vc{q} = - \kappa(\vt) \Grad \vt,
\eF
and $\sigma$ stands for the entropy production rate specified below. The system of equations (\ref{i1} - \ref{i3}), with the constitutive relations (\ref{i4}), (\ref{i5}) is called \emph{Navier-Stokes-Fourier system}.
Equations (\ref{i1} - \ref{i4}) are supplemented with \emph{conservative} boundary conditions, say,
\bFormula{i6}
\vu|_{\partial \Omega} = \vc{q} \cdot \vc{n}|_{\partial \Omega} = 0.
\eF
A concept of \emph{weak solution} to the Navier-Stokes-Fourier
system (\ref{i1} - \ref{i6}) based on Second law of thermodynamics
was introduced in \cite{DUFE2}. The weak solutions satisfy the
field equations (\ref{i1} - \ref{i3}) in the sense of
distributions, where the entropy production $\sigma$ is a
non-negative measure, \bFormula{i7} \sigma \geq \frac{1}{\vt}
\left( \tn{S}(\vt, \Grad \vu) : \Grad \vu - \frac{\vc{q}(\vt,
\Grad \vt) \cdot \Grad \vt}{\vt} \right). \eF In order to
compensate for the lack of information resulting from the
inequality sign in (\ref{i7}), the resulting system is
supplemented by the total energy balance, \bFormula{i8} \frac{{\rm
d}}{{\rm d}t} \intO{ \left( \frac{1}{2} \vr |\vu|^2 + \vr e(\vr,
\vt) \right) } = 0, \eF where $e= e(\vr, \vt)$ is the (specific)
internal energy. Under these circumstances, it can be shown (see
\cite[Chapter 2]{FENO6}) that any weak solution of (\ref{i1}) that
is sufficiently \emph{smooth} satisfies, instead of (\ref{i7}),
the standard relation \bFormula{i9} \sigma = \frac{1}{\vt} \left(
\tn{S}(\vt, \Grad \vu) : \Grad \vu - \frac{\vc{q}(\vt, \Grad \vt)
\cdot \Grad \vt}{\vt} \right). \eF
Despite the fact that the framework of the weak solutions is sufficiently robust to provide:
\begin{itemize}
\item
a general \emph{existence theory} of global-in-time solutions under certain restrictions imposed on the
constitutive relations, see \cite[Chapter 3]{FENO6};
\item
quite satisfactory and complete description of the long-time behavior of solutions, see \cite{FeiPr};
\item
rigorous justification of certain singular limits, see \cite{FENO6};
\end{itemize}
the class of weak solutions satisfying (\ref{i7}), (\ref{i8})
instead of (\ref{i9}) is probably too large to give rise to a
unique solution of the associated initial-boundary value problem.
The main issue seems to be the ambiguity of possible continuation
of solutions in the hypothetical presence of the so-called vacuum
zones (where $\vr = 0$) in combination with the possibility of
concentrations in the entropy production rate $\sigma$. On the
other hand, as we show in the present paper, any weak solution of
the Navier-Stokes-Fourier system coincides with a strong solution
as long as the latter exists. This is another piece of evidence
that the weak solutions, in the sense specified above, represent a
suitable extension of classical smooth solutions beyond their
(hypothetical) ``blow-up'' time.
Our approach is based on the method of \emph{relative entropy}
(cf. Carrillo et al. \cite{CaJuMaToUn}, Dafermos \cite{Daf4}, Saint-Raymond \cite{SaRay}), represented in the present context by the quantity
\bFormula{i10}
\mathcal{E}(\vr, \vt \ | \tilde \vr , \tilde \vt ) =
H_{\tilde \vt} (\vr, \vt) - \partial_\vr H_{\tilde \vt} (\tilde \vr, \tilde \vt) (\vr - \tilde \vr) - H_{\tilde \vt} (\tilde \vr, \tilde \vt),
\eF
where $H_{\tilde \vt}(\vr, \vt)$ is a thermodynamic potential termed \emph{ballistic free energy}
\bFormula{i11}
H_{\tilde \vt} (\vr, \vt) = \vr e(\vr, \vt) - \tilde \vt \vr s(\vr , \vt),
\eF
introduced by Gibbs and discussed more recently by Ericksen \cite{Eri}.
We assume that the thermodynamic functions $p$, $e$, and $s$ are interrelated through Gibbs' equation
\bFormula{i12}
\vt D s(\vr, \vt) = D e(\vr, \vt) + p(\vr, \vt) D \left( \frac{1}{\vr} \right).
\eF
The subsequent analysis leans essentially on \emph{thermodynamic stability} of the fluid system expressed through
\bFormula{i13}
\frac{\partial p(\vr, \vt)}{\partial \vr} > 0,\ \frac{\partial e(\vr, \vt)}{\partial \vt} > 0
\ \mbox{for all} \ \vr, \vt > 0.
\eF
In terms of the function $H_{\tilde \vt}$, thermodynamic stability implies that
\bFormula{i14}
\vr \mapsto H_{\tilde \vt} (\vr, \tilde \vt) \ \mbox{is strictly convex,}
\eF
while
\bFormula{i15}
\vt \mapsto H_{\tilde \vt}(\vr, \vt) \ \mbox{attains its global minimum at}\ \vt = \tilde \vt.
\eF
The above relations reflect \emph{stability} of the equilibrium solutions to the Navier-Stokes-Fourier system
(see Bechtel, Rooney, and Forest \cite{BEROFO}) and play a crucial role in the study of the long-time behavior
of solutions, see \cite[Chapters 5,6]{FeiPr}. Moreover, properties (\ref{i14}), (\ref{i5}) yield the necessary uniform bounds in
singular limits with ill-prepared initial data, see \cite[Chapter 2]{FENO6}. Last but not least, as we will see below,
the relations (\ref{i14}), (\ref{i15}) represent the key ingredient in the proof of weak-strong uniqueness.
The weak-strong uniqueness property for the standard
\emph{incompressible} Navier-Stokes system was established in
seminal papers by Prodi \cite{PR} and Serrin \cite{Serrin}. The
situation is a bit more delicate in the case of a
\emph{compressible} fluid. Germain \cite{Ger} proves weak-strong
uniqueness for the isentropic Navier-Stokes system, unfortunately,
in the class of ``more regular'' weak solutions which existence is
not known, see also Desjardin \cite{DES2} for the previous
results in this direction. The problem is finally settled in \cite{FeJiNo}
(see also \cite{FENOSU}), where the weak-strong uniqueness is
established in the class of the weak solutions to the barotropic
Navier-Stokes system satisfying the energy inequality.
Our method leans on a kind of \emph{total dissipation balance} stated in terms of the quantity
\[
I = \intO{ \left( \frac{1}{2} \vr |\vu - \tilde \vu |^2 + \mathcal{E}(\vr, \vt | \tilde \vr, \tilde \vt ) \right)} ,
\]
where $\{ \vr, \vt, \vu \}$ is a weak solution to the Navier-Stokes-Fourier system and
$\{ \tilde \vr, \tilde \vt, \tilde \vu \}$ is a hypothetical classical solution emanating from the same initial data. We use a Gronwall type argument to show that $I(t) = 0$ provided
$I(0) = 0$ as long as the classical solution exists. The proof is carried over under the list of structural hypotheses
that are identical to those providing the existence of global-in-time solutions. Possible relaxations are discussed
in the concluding part of the paper.
The paper is organized as follows. In Section \ref{m}, we recall the definition of the weak solutions to the Navier-Stokes-Fourier system, introduce the hypotheses imposed on the
constitutive relations, and state our main result. In Section \ref{d}, we deduce a version of relative entropy inequality by taking $\tilde \vr = r$, $\tilde \vt = \Theta$, $\tilde \vu = \vc{U}$,
where $r$, $\Theta$, and $\vc{U}$ are arbitrary smooth functions satisfying the relevant boundary conditions. Finally, in Section \ref{w}, we establish the weak-strong uniqueness property. Possible extensions and further applications of the method are discussed in Section \ref{c}.
\section{Main result}
\label{m}
Motivated by the existence theory developed in \cite[Chapter 3]{FENO6}, we assume that the pressure $p = p(\vr, \vt)$ can be written in the form
\bFormula{m1}
p(\vr, \vt) = \vt^{5/2} P \left( \frac{\vr}{\vt^{3/2}} \right) + \frac{a}{3} \vt^4, \
a > 0,
\eF
where
\bFormula{m2}
P \in C^1[0, \infty), \ P(0) = 0, \ P'(Z) > 0 \ \mbox{for all}\ Z \geq 0.
\eF
In agreement with Gibbs' relation (\ref{i12}), the (specific) internal energy can be taken as
\bFormula{m3}
e(\vr, \vt) = \frac{3}{2} \frac{\vt^{5/2}}{\vr}P \left( \frac{\vr}{\vt^{3/2}} \right) +
a \frac{\vt^4}{\vr}.
\eF
Furthermore, by virtue of the second inequality in thermodynamic stability hypothesis (\ref{i13}), we have
\bFormula{m4}
0 < \frac{\frac{5}{3} P(Z) - P'(Z) Z }{Z} < c \ \mbox{for all}\ Z > 0.
\eF
Relation (\ref{m4}) implies that the function $Z \mapsto P(Z) / Z^{5/3}$ is decreasing, and we suppose that
\bFormula{m5}
\lim_{Z \to \infty} \frac{P(Z)}{Z^{5/3}} = P_\infty > 0.
\eF
Finally, the formula for (specific) entropy reads
\bFormula{m6}
s(\vr, \vt) = S \left( \frac{\vr}{\vt^{3/2}} \right) + \frac{4a}{3} \frac{\vt^3}{\vr},
\eF
where, in accordance with Third law of thermodynamics,
\bFormula{m7}
S'(Z) = - \frac{3}{2} \frac{ \frac{5}{3} P(Z) - P'(Z) Z }{Z^2} < 0,\
\lim_{ Z \to \infty } S(Z) = 0.
\eF
The reader may consult Eliezer, Ghatak, and Hora \cite{EGH} and \cite[Chapter 3]{FENO6}
for the physical background and further discussion concerning the structural hypotheses
(\ref{m1} - \ref{m5}).
For the sake of simplicity and clarity of presentation, we take the transport coefficients in the form
\bFormula{m8}
\mu (\vt) = \mu_0 + \mu_1 \vt ,\ \mu_0, \mu_1 > 0, \ \eta \equiv 0,
\eF
\bFormula{m9}
\kappa (\vt) = \kappa_0 + \kappa_2 \vt^2 + \kappa_3 \vt^3 ,\ \kappa_i > 0, \ i=0,2,3,
\eF
cf. \cite[Chapter 3]{FENO6}.
\subsection{Weak solutions to the Navier-Stokes-Fourier system}
\label{ws}
Let $\Omega \subset R^3$ be a bounded Lipschitz domain. We say
that a trio $\{ \vr, \vt, \vu \}$ is a \emph{weak solution} to the
Navier-Stokes-Fourier system (\ref{i1} - \ref{i8}) emanating from
the initial data \bFormula{m10} \vr(0, \cdot) = \vr_0, \ \vr \vu
(0, \cdot) = \vr_0 \vu_0,\ \vr s(\vr, \vt)(0, \cdot) = \vr_0
s(\vr_0, \vt_0),\quad{ \vr_0\ge 0,\;\vt_0>0} \eF if:
\begin{itemize}
\item the density and the absolute temperature satisfy $\vr(t,x) \geq 0$, $\vt(t,x) > 0$ for a.a. $(t,x) \in (0,T) \times \Omega$, $\vr
\in C_{\rm weak}([0,T]; L^{5/3})$, $\vr \vu \in C_{\rm weak}([0,T]; L^{5/4}(\Omega;R^3))$,
$\vt \in L^\infty(0,T; L^4(\Omega)) \cap L^2(0,T; W^{1,2}(\Omega))$, and $\vu \in L^2(0,T; W^{1,2}_0(\Omega;R^3))$;
\item equation (\ref{i1}) is replaced by a family of integral identities
\bFormula{m11} \intO{ \vr (\tau, \cdot) \varphi (\tau, \cdot)}
-\intO{ \vr_0 \varphi (0, \cdot) } = \int_0^\tau \intO{ \Big( \vr
\partial_t \varphi + \vr \vu \cdot \Grad \varphi \Big) } \ \dt \eF
for any $\varphi \in C^1([0,T] \times \Ov{\Omega})$, and any $\tau
\in [0,T]$;
\item momentum equation (\ref{i2}) is satisfied in the sense of distributions, specifically,
\bFormula{m12}
\intO{ \vr \vu (\tau, \cdot) \cdot \varphi (\tau, \cdot) } -
\intO{ \vr_0 \vu_0 \cdot \varphi (0, \cdot) }
\eF
\[
\int_0^\tau \intO{ \Big( \vr \vu \cdot \partial_t \varphi + \vr \vu \otimes \vu : \Grad \varphi + p(\vr, \vt) \Div \varphi - \tn{S}(\vt, \Grad \vu) : \Grad \varphi \Big) } \ \dt
\]
for any $\varphi \in C^1([0,T] \times \Ov{\Omega}; R^3)$, $\varphi|_{\partial \Omega} = 0$,
and any $\tau \in [0,T]$;
\item the entropy balance (\ref{i3}), (\ref{i7}) is replaced by a family of integral inequalities
\bFormula{m13}
\intO{ \vr_0 s(\vr_0, \vt_0) \varphi (0, \cdot) } -
\intO{ \vr s(\vr, \vt) (\tau, \cdot) \varphi (\tau, \cdot) }
\eF
\[
+
\int_0^\tau \intO{ \frac{ \varphi }{\vt} \left(
\tn{S}(\vt, \Grad \vu) : \Grad \vu - \frac{\vc{q}(\vt, \Grad \vt) \cdot \Grad \vt }{\vt}
\right) } \ \dt
\]
\[
\leq - \int_0^\tau \intO{ \left( \vr s(\vr, \vt) \partial_t \varphi + \vr s(\vr, \vt) \vu \cdot \Grad \varphi + \frac{ \vc{q}(\vt, \Grad \vt) \cdot \Grad \varphi }{\vt} \right)
} \ \dt
\]
for any $\varphi \in C^1([0,T] \times \Ov{\Omega})$, $\varphi \geq 0$, and a.a. $\tau \in [0,T]$;
\item the total energy is conserved:
\bFormula{m14}
\intO{ \left( \frac{1}{2} \vr |\vu|^2 + \vr e(\vr, \vt) \right)(\tau, \cdot) } =
\intO{ \left( \frac{1}{2} \vr_0 |\vu_0|^2 + \vr_0 e(\vr_0, \vt_0) \right) }
\eF
for a.a. $\tau \in [0,T]$.
\end{itemize}
The \emph{existence} of global-in-time weak solutions to the Navier-Stokes-Fourier system was established in \cite[Chapter 3, Theorem 3.1]{FENO6}.
\subsection{Main result}
We say that $\{ \tilde \vr, \tilde \vt, \tilde \vu \}$ is a
classical (strong) solution to the Navier-Stokes-Fourier system in
$(0,T) \times \Omega$ if
\bFormula{2.15} \tilde \vr \in C^1([0,T] \times \Ov{\Omega}),\
\tilde \vt,\ \partial_t \tilde \vt, \ \nabla^2 \tilde \vt \in
C([0,T] \times \Ov{\Omega}), \ \tilde \vu, \
\partial_t \tilde \vu,\ \nabla^2 \tilde \vu \in C([0,T] \times
\Ov{\Omega};R^3), \eF
\[
\tilde \vr (t,x) \geq \underline{\vr} > 0 , \ \tilde \vt (t,x) \geq \underline{\vt} > 0
\ \mbox{for all}\ (t,x),
\]
and $\tilde \vr$, $\tilde \vt$, $\tilde \vu$ satisfy equations
(\ref{i1} - \ref{i3}), (\ref{i9}), together with the boundary
conditions (\ref{i6}).
{Observe that hypothesis (\ref{2.15}) implies the following regularity properties of the initial data:
\bFormula{2.16} \vr(0)=\vr_0\in
C^1(\overline\Omega),\quad\vr_0\ge\underline\vr>0, \eF
\[
\vt(0)=\vt_0\in C^2(\overline\Omega),\quad\vt_0\ge\underline\vt>0,
\]
\[
\vu(0)=\vu_0\in C^2(\overline\Omega).
\]
}
We are ready to state the main result of this paper.
\bTheorem{m1} Let $\Omega \subset R^3$ be a bounded Lipschitz
domain. Suppose that the thermodynamic functions $p$, $e$, $s$
satisfy hypotheses (\ref{m1} - \ref{m7}), and that the transport
coefficients $\mu$, $\eta$, and $\kappa$ obey (\ref{m8}),
(\ref{m9}). Let $\{ \vr, \vt, \vu \}$ be a weak solution of the
Navier-Stokes-Fourier system in $(0,T) \times \Omega$ in the sense
specified in Section \ref{ws}, and let $\{ \tilde \vr, \tilde \vt,
\tilde \vu \}$ be a strong solution emanating from the same
initial data
{(\ref{2.16}).
}
Then
\[
\vr \equiv \tilde \vr, \ \vt \equiv \tilde \vt,\ \vu \equiv \tilde \vu.
\]
\eT
The rest of the paper is devoted to the proof of Theorem \ref{Tm1}. Possible generalizations are discussed in Section \ref{c}.
\section{Relative entropy inequality}
\label{d}
We deduce a relative entropy inequality satisfied by \emph{any} weak solution to the Navier-Stokes-Fourier system. To this end, consider a trio $\{ r , \Theta, \vc{U} \}$ of
smooth functions, $r$ and $\Theta$ bounded below away from zero in $[0,T] \times \Omega$, and
$\vc{U}|_{\partial \Omega} = 0$.
Taking $\varphi = \frac{1}{2} |\vc{U}|^2$ as a test function in (\ref{m11}), we get
\bFormula{d1}
\intO{ \frac{1}{2} \vr |\vc{U}|^2 (\tau, \cdot) } = \intO{ \frac{1}{2} \vr_0 |\vc{U}(0, \cdot)|^2 } +
\int_0^\tau \intO{ \Big( \vr \vc{U} \cdot \partial_t \vc{U} + \vr \vu \cdot \Grad \vc{U} \cdot \vc{U} \Big) } \ \dt.
\eF
Similarly, the choice $\varphi = \vc{U}$ in (\ref{m12}) gives rise to
\bFormula{d2}
\intO{ \vr \vu \cdot \vc{U} (\tau, \cdot) }
- \intO{ \vr_0 \vu_0 \cdot \vc{U}(0, \cdot) }
\eF
\[
= \int_0^\tau \intO{ \Big( \vr \vu \cdot \partial_t \vc{U} + \vr \vu \otimes \vu : \Grad \vc{U} +
p(\vr, \vt) \Div \vc{U} - \tn{S}(\vt, \Grad \vu) : \Grad \vc{U} \Big) } \ \dt.
\]
Combining relations (\ref{d1}), (\ref{d2}) with the total energy balance (\ref{m14}) we may infer that
\bFormula{d3}
\intO{ \left( \frac{1}{2} \vr | \vu - \vc{U} |^2 + \vr e(\vr, \vt) \right)(\tau, \cdot) } =
\intO{ \left( \frac{1}{2} \vr_0 |\vu_0 - \vc{U}(0, \cdot) |^2 + \vr_0 e(\vr_0, \vt_0) \right) }
\eF
\[
+ \int_0^\tau \intO{ \left( \Big( \vr \partial_t \vc{U} + \vr \vu \cdot \Grad \vc{U} \Big) \cdot (\vc{U} - \vu) - p(\vr, \vt) \Div \vc{U} + \tn{S}(\vt, \Grad \vu) : \Grad \vc{U} \right) } \ \dt.
\]
Now, take $\varphi = \Theta > 0$ as a test function in the entropy inequality (\ref{m13})
to obtain
\bFormula{d4}
\intO{ \vr_0 s(\vr_0, \vt_0) \Theta(0, \cdot) } -
\intO{ \vr s(\vr, \vt) \Theta (\tau, \cdot) }
\eF
\[
+
\int_0^\tau \intO{ \frac{\Theta}{\vt} \left(
\tn{S}(\vt, \Grad \vu): \Grad \vu - \frac{ \vc{q}(\vt, \Grad \vt) \cdot \Grad \vt }{\vt} \right) } \ \dt
\]
\[
\leq - \int_0^\tau \intO{ \left( \vr s(\vr, \vt) \partial_t \Theta + \vr s(\vr, \vt) \vu \cdot \Grad \Theta + \frac{ \vc{q}(\vt, \Grad \vt) }{\vt} \cdot \Grad \Theta \right) } \ \dt.
\]
Thus, the sum of (\ref{d3}), (\ref{d4}) reads
\bFormula{d5}
\intO{ \left( \frac{1}{2} \vr | \vu - \vc{U} |^2 + \vr e(\vr, \vt) -
\Theta \vr s(\vr, \vt) \right)(\tau, \cdot) }
\eF
\[
+
\int_0^\tau \intO{ \frac{\Theta}{\vt} \left(
\tn{S}(\vt, \Grad \vu): \Grad \vu - \frac{ \vc{q}(\vt, \Grad \vt) \cdot \Grad \vt }{\vt} \right) } \ \dt
\]
\[
=
\intO{ \left( \frac{1}{2} \vr_0 |\vu_0 - \vc{U}(0, \cdot) |^2 + \vr_0 e(\vr_0, \vt_0)
- \Theta(0, \cdot) \vr_0 s(\vr_0, \vt_0) \right) }
\]
\[
+ \int_0^\tau \intO{ \left( \Big( \vr \partial_t \vc{U} + \vr \vu \cdot \Grad \vc{U} \Big) \cdot (\vc{U} - \vu) - p(\vr, \vt) \Div \vc{U} + \tn{S}(\vt, \Grad \vu) : \Grad \vc{U} \right) } \ \dt
\]
\[
- \int_0^\tau \intO{ \left( \vr s(\vr, \vt) \partial_t \Theta + \vr s(\vr, \vt) \vu \cdot \Grad \Theta + \frac{ \vc{q}(\vt, \Grad \vt) }{\vt} \cdot \Grad \Theta \right) } \ \dt .
\]
Next, we take $\varphi = \partial_\vr H_\Theta(r, \Theta)$ as a test function in (\ref{m11}) to deduce that
\[
\intO{ \vr \partial_\vr H_\Theta (r, \Theta) (\tau, \cdot)} =
\intO{ \vr_0 \partial_\vr H_{\Theta(0, \cdot)} (r(0, \cdot), \Theta(0, \cdot) }
\]
\[
+ \int_0^\tau \intO{ \left(
\vr \partial_t \Big( \partial_\vr H_\Theta (r, \Theta) \Big) + \vr \vu \cdot
\Grad \Big( \partial_\vr H_\Theta (r, \Theta) \Big) \right) } \ \dt,
\]
which, combined with (\ref{d5}), gives rise to
\bFormula{d6}
\intO{ \left( \frac{1}{2} \vr | \vu - \vc{U}|^2 + H_\Theta (\vr, \vt) -
\partial_\vr (H_\Theta)(r, \Theta)(\vr - r) - H_\Theta (r, \Theta) \right)(\tau, \cdot)} +
\eF
\[
\int_0^\tau \intO{ \frac{\Theta}{\vt} \left(
\tn{S}(\vt, \Grad \vu): \Grad \vu - \frac{ \vc{q}(\vt, \Grad \vt) \cdot \Grad \vt }{\vt} \right) } \ \dt
\]
\[
\leq \intO{ \frac{1}{2} \vr_0 | \vu_0 - \vc{U}(0, \cdot)|^2 }
\]
\[
+ \intO{ \left( H_{\Theta(0, \cdot)} (\vr_0, \vt_0) -
\partial_\vr (H_{\Theta(0, \cdot)})(r(0, \cdot), \Theta(0, \cdot))(\vr_0 - r(0, \cdot)) - H_{\Theta(0, \cdot)} (r(0, \cdot), \Theta(0, \cdot)) \right)}
\]
\[
+ \int_0^\tau \intO{ \left( \Big( \vr \partial_t \vc{U} + \vr \vu \cdot \Grad \vc{U} \Big) \cdot (\vc{U} - \vu) - p(\vr, \vt) \Div \vc{U} + \tn{S}(\vt, \Grad \vu) : \Grad \vc{U} \right) } \ \dt
\]
\[
- \int_0^\tau \intO{ \left( \vr s(\vr, \vt) \partial_t \Theta + \vr s(\vr, \vt) \vu \cdot \Grad \Theta + \frac{ \vc{q}(\vt, \Grad \vt) }{\vt} \cdot \Grad \Theta \right) } \ \dt .
\]
\[
- \int_0^\tau \intO{ \left(
\vr \partial_t \Big( \partial_\vr H_\Theta (r, \Theta) \Big) + \vr \vu \cdot
\Grad \Big( \partial_\vr H_\Theta (r, \Theta) \Big) \right) } \ \dt
\]
\[
+ \int_0^\tau \intO{ \partial_t \Big( r \partial_\vr (H_\Theta) (r, \Theta) -
H_\Theta (r, \Theta) \Big) } \ \dt.
\]
Furthermore, seeing that
\[
\partial_y \left( \partial_\vr H_\Theta (r, \Theta) \right) = - s(r, \Theta) \partial_y \Theta - r \partial_\vr (r, \Theta) \partial_y \Theta + \partial^2_{\vr,\vr} H_\Theta (r, \Theta) \partial_y \vr + \partial^2_{\vr, \vt} H_\Theta (r, \Theta) \partial_y \Theta
\]
for $y=t,x$,
we may rewrite (\ref{d6}) in the form
\bFormula{d7}
\intO{ \left( \frac{1}{2} \vr | \vu - \vc{U}|^2 + H_\Theta (\vr, \vt) -
\partial_\vr (H_\Theta)(r, \Theta)(\vr - r) - H_\Theta (r, \Theta) \right)(\tau, \cdot)} +
\eF
\[
\int_0^\tau \intO{ \frac{\Theta}{\vt} \left(
\tn{S}(\vt, \Grad \vu): \Grad \vu - \frac{ \vc{q}(\vt, \Grad \vt) \cdot \Grad \vt }{\vt} \right) } \ \dt
\]
\[
\leq \intO{ \frac{1}{2} \vr_0 | \vu_0 - \vc{U}(0, \cdot)|^2 }
\]
\[
+ \intO{ \left( H_{\Theta(0, \cdot)} (\vr_0, \vt_0) -
\partial_\vr (H_{\Theta(0, \cdot)})(r(0, \cdot), \Theta(0, \cdot))(\vr_0 - r(0, \cdot)) - H_{\Theta(0, \cdot)} (r(0, \cdot), \Theta(0, \cdot)) \right)}
\]
\[
+ \int_0^\tau \intO{ \left( \Big( \vr \partial_t \vc{U} + \vr \vu \cdot \Grad \vc{U} \Big) \cdot (\vc{U} - \vu) - p(\vr, \vt) \Div \vc{U} + \tn{S}(\vt, \Grad \vu) : \Grad \vc{U} \right) } \ \dt
\]
\[
- \int_0^\tau \intO{ \left( \vr \Big( s (\vr, \vt) - s(r, \Theta) \Big) \partial_t \Theta
+ \vr \Big( s(\vr, \vt) - s(r, \Theta) \Big) \vu \cdot \Grad \Theta +
\frac{ \vc{q}(\vt, \Grad \vt) }{\vt} \cdot \Grad \Theta \right) } \ \dt .
\]
\[
+ \int_0^\tau \intO{ \vr \Big( r \partial_\vr s(r, \Theta) \partial_t \Theta +
r \partial_\vr s(r, \Theta) \vu \cdot \Grad \Theta \Big) } \ \dt
\]
\[
- \int_0^\tau \intO{
\vr \Big( \partial^2_{\vr, \vr} (H_\Theta) (r, \Theta) \partial_t r +
\partial^2_{\vr, \vt} (H_\Theta) (r, \Theta) \partial_t \Theta \Big) } \ \dt
\]
\[
- \int_0^\tau \intO{ \vr \vu \cdot
\Big( \partial^2_{\vr, \vr}( H_\Theta) (r, \Theta) \Grad r +
\partial^2_{\vr, \vt} (H_\Theta) (r, \Theta) \Grad \Theta \Big) \Big) } \ \dt
\]
\[
+ \int_0^\tau \intO{ \partial_t \Big( r \partial_\vr (H_\Theta) (r, \Theta) -
H_\Theta (r, \Theta) \Big) } \ \dt.
\]
In order to simplify (\ref{d7}), we recall several useful identities that follow directly from Gibbs' relation (\ref{i12}):
\[
\partial^2_{\vr, \vr} (H_\Theta) (r, \Theta) = \frac{1}{r} \partial_\vr p (r, \Theta),
\]
\bFormula{sim}
r \partial_\vr s(r, \Theta) = - \frac{1}{r} \partial_\vt p (r, \Theta),
\eF
\[
\partial^2_{\vr, \vt} (H_\Theta) (r, \Theta) = \partial_\vr \left( \vr ( \vt - \Theta )
\partial_\vt s \right) (r, \Theta) = (\vt - \Theta) \partial_\vr \Big( \vr \partial_\vt s(\vr, \vt) \Big) (r, \Theta) = 0,
\]
and
\[
r \partial_\vr (H_\Theta) (r, \Theta) - H_\Theta(r, \Theta) = p(r, \Theta).
\]
Thus relation (\ref{d7}) can be finally written in a more concise form
\bFormula{d8}
\intO{ \left( \frac{1}{2} \vr | \vu - \vc{U}|^2 + \mathcal{E}(\vr, \vt | r, \Theta) \right)(\tau, \cdot)} +
\int_0^\tau \intO{ \frac{\Theta}{\vt} \left(
\tn{S}(\vt, \Grad \vu): \Grad \vu - \frac{ \vc{q}(\vt, \Grad \vt) \cdot \Grad \vt }{\vt} \right) } \ \dt
\eF
\[
\leq \intO{ \left( \frac{1}{2} \vr_0 | \vu_0 - \vc{U}(0, \cdot)|^2 +
\mathcal{E} (\vr_0, \vt_0 | r(0, \cdot), \Theta (0, \cdot)) \right)}
\]
\[
\int_0^\tau \intO{ \vr (\vu - \vc{U}) \cdot \Grad \vc{U} \cdot (\vc{U} - \vu)} \ \dt
+ \int_0^\tau \intO{ \vr \Big( s(\vr, \vt) - s(r, \Theta) \Big) \Big( \vc{U} - \vu \Big)
\cdot \Grad \Theta } \ \dt
\]
\[
+ \int_0^\tau \intO{ \left( \vr \Big( \partial_t \vc{U} + \vc{U} \cdot \Grad \vc{U} \Big) \cdot (\vc{U} - \vu) - p(\vr, \vt) \Div \vc{U} + \tn{S}(\vt, \Grad \vu) : \Grad \vc{U} \right) } \ \dt
\]
\[
- \int_0^\tau \intO{ \left( \vr \Big( s (\vr, \vt) - s(r, \Theta) \Big) \partial_t \Theta
+ \vr \Big( s(\vr, \vt) - s(r, \Theta) \Big) \vc{U} \cdot \Grad \Theta +
\frac{ \vc{q}(\vt, \Grad \vt) }{\vt} \cdot \Grad \Theta \right) } \ \dt
\]
\[
+ \int_0^\tau \intO{ \left( \left( 1 - \frac{\vr}{r} \right) \partial_t p(r, \Theta) -
\frac{\vr}{r} \vu \cdot \Grad p(r, \Theta) \right) } \ \dt,
\]
where $\mathcal{E}$ was introduced in (\ref{i10}).
Formula (\ref{d8}) represents a kind of \emph{relative entropy inequality} in the spirit of
\cite{FeJiNo}, \cite{FENOSU}. Note that it is satisfied for \emph{any} trio of smooth functions
$\{r, \Theta, \vc{U}\}$ provided $\vc{U}$ vanishes on the boundary $\partial \Omega$, meaning
$\vc{U}$ is an admissible test function in the weak formulation of the momentum equation (\ref{m12}). Similar result can be obtained for other kinds of boundary conditions. The requirement on smoothness of $r$, $\Theta$, and $\vc{U}$ can be relaxed given the specific
integrability properties of the weak solution $\vr$, $\vt$, $\vu$. It is also worth-noting that (\ref{d8}) holds provided $p$, $e$, and $s$ satisfy only Gibbs' equation, the structural restrictions introduced in Section \ref{m} are not needed at this step.
\section{Weak-strong uniqueness}
\label{w}
In this section we finish the proof of Theorem \ref{Tm1} by applying the relative entropy inequality (\ref{d8}) to $r= \tilde \vr$, $\Theta = \tilde \vt$, and $\vc{U} = \tilde \vu$,
where $\{ \tilde \vr, \tilde \vt, \tilde \vu \}$ is a classical (smooth) solution of the Navier-Stokes-Fourier system such that
\[
\tilde \vr (0, \cdot) = \vr_0 , \ \tilde \vu (0, \cdot) = \vu_0,\ \tilde \vt (0, \cdot) =
\vt_0.
\]
Accordingly, the integrals depending on the initial values on the right-hand side of (\ref{d8}) vanish, and we apply a Gronwall type argument to deduce the desired result, namely,
\[
\vr \equiv \tilde \vr ,\ \vt \equiv \tilde \vt, \ \mbox{and}\ \vu \equiv \tilde \vu.
\]
Here, the hypothesis of thermodynamic stability formulated in (\ref{i13}) will play a crucial role.
\subsection{Preliminaries, notation}
Following \cite[Chapters 4,5]{FENO6} we introduce \emph{essential} and \emph{residual} component of each quantity appearing in (\ref{d8}). To begin, we choose positive constants
$\underline{\vr}$, $\Ov{\vr}$, $\underline{\vt}$, $\Ov{\vt}$ in such a way that
\[
0 < \underline{\vr} \leq \frac{1}{2} \min_{(t,x) \in [0,T] \times \Ov{\Omega}} \tilde \vr(t,x) \leq 2 \max_{(t,x) \in [0,T] \times \Ov{\Omega}} \tilde \vr (t,x) \leq
\Ov{\vr},
\]
\[
0 < \underline{\vt} \leq \frac{1}{2} \min_{(t,x) \in [0,T] \times \Ov{\Omega}} \tilde \vt(t,x) \leq 2 \max_{(t,x) \in [0,T] \times \Ov{\Omega}} \tilde \vt (t,x) \leq
\Ov{\vt}.
\]
In can be shown, as a consequence of the hypothesis of thermodynamic stability (\ref{i13}),
or, more specifically, of (\ref{i14}), (\ref{i15}), that
\bFormula{w1}
\mathcal{E} (\vr, \vt | \tilde \vr, \tilde \vt ) \geq c \left\{
\begin{array}{l} |\vr - \tilde \vr |^2 + |\vt - \tilde \vt |^2
\ \mbox{if} \ (\vr, \vt) \in [\underline{\vr}, \overline{\vr}]
\times [\underline{\vt}, \Ov{\vt}] \\ \\
1 + |\vr s(\vr, \vt) | + \vr e(\vr, \vt) \ \mbox{otherwise,}
\end{array}
\right.
\eF
whenever $[\tilde \vr, \tilde \vt] \in [\underline{\vr}, \overline{\vr}]
\times [\underline{\vt}, \Ov{\vt}]$, where the constant $c$ depends only on
$\underline{\vr}$, $\Ov{\vr}$, $\underline{\vt}$, $\Ov{\vt}$ and the structural properties of the
thermodynamic functions $e$, $s$, see \cite[Chapter 3, Proposition 3.2]{FENO6}.
Now, each measurable function $h$ can be written as
\[
h = h_{\rm ess} + h_{\rm res},
\]
where
\[
h_{\rm ess}(t,x) = \left\{ \begin{array}{l} h(t,x) \ \mbox{if}\ (\vr(t,x), \vt(t,x)) \in [\underline{\vr}, \overline{\vr}]
\times [\underline{\vt}, \Ov{\vt}] \\ \\ 0 \ \mbox{otherwise}
\end{array} \right. ,\ h_{\rm res} = h - h_{\rm ess}.
\]
\subsection{Relative entropy balance}
Taking $r = \tilde \vr$, $\Theta = \tilde \vt$, $\vc{U} = \tilde \vu$ in (\ref{d8}) and using the fact that the initial values coincide, we obtain
\bFormula{w2}
\intO{ \left( \frac{1}{2} \vr | \vu - \tilde \vu |^2 + \mathcal{E}(\vr, \vt | \tilde \vr, \tilde \vt ) \right)(\tau, \cdot)} +
\int_0^\tau \intO{ \frac{\tilde \vt}{\vt} \left(
\tn{S}(\vt, \Grad \vu): \Grad \vu - \frac{ \vc{q}(\vt, \Grad \vt) \cdot \Grad \vt }{\vt} \right) } \ \dt
\eF
\[
\leq \int_0^\tau \intO{ \vr |\vu - \tilde \vu|^2 |\Grad \tilde \vu| } \ \dt
+ \int_0^\tau \intO{ \vr \Big( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \Big) \Big( \tilde \vu - \vu \Big)
\cdot \Grad \tilde \vt } \ \dt
\]
\[
+ \int_0^\tau \intO{ \left( \vr \Big( \partial_t \tilde \vu + \tilde \vu \cdot \Grad \tilde \vu \Big) \cdot (\tilde \vu - \vu) - p(\vr, \vt) \Div \tilde \vu + \tn{S}(\vt, \Grad \vu) : \Grad \tilde \vu \right) } \ \dt
\]
\[
- \int_0^\tau \intO{ \left( \vr \Big( s (\vr, \vt) - s(\tilde \vr, \tilde \vt) \Big) \partial_t \tilde \vt
+ \vr \Big( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \Big) \tilde \vu \cdot \Grad \tilde \vt + \frac{ \vc{q}(\vt, \Grad \vt) }{\vt} \cdot \Grad \tilde \vt \right) } \ \dt
\]
\[
+ \int_0^\tau \intO{ \left( \left( 1 - \frac{\vr}{\tilde \vr} \right) \partial_t p(\tilde \vr, \tilde \vt) -
\frac{\vr}{\tilde \vr} \vu \cdot \Grad p(\tilde \vr, \tilde \vt) \right) } \ \dt.
\]
In order to handle the integrals on the right-hand side of (\ref{w2}), we proceed by several steps:
\medskip
{\bf Step 1:}
We have
\bFormula{w3}
\intO{ \vr |\vu - \tilde \vu|^2 |\Grad \tilde \vu| } \ \dt \leq
2 \| \Grad \tilde \vu \|_{L^\infty(\Omega; R^{3 \times 3} )}
\intO{ \frac{1}{2} \vr | \vu - \tilde \vu |^2 }.
\eF
\medskip
{\bf Step 2:}
\[
\left| \intO{ \vr \Big( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \Big) \Big( \tilde \vu - \vu \Big)
\cdot \Grad \tilde \vt } \right|
\]
\[
\| \Grad \tilde \vt \|_{L^\infty(\Omega;R^3)} \left[
2 \Ov{\vr} \intO{ \left| \left[ s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \right]_{\rm ess} \right|
|\vu - \tilde \vu | } + \intO{ \left| \left[ \vr \left( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \right) \right]_{\rm res} \right| |\vu - \tilde \vu| }
\right],
\]
where, by virtue of (\ref{w1}),
\bFormula{w4}
\intO{ \left| \left[ s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \right]_{\rm ess} \right|
|\vu - \tilde \vu | } \leq \delta \| \vu - \tilde \vu \|^2_{L^2(\Omega; R^3)} +
c(\delta) \intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde \vt ) }
\eF
for any $\delta > 0$.
Similarly, by interpolation inequality,
\[
\intO{ \left| \left[ \vr \left( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \right) \right]_{\rm res} \right| |\vu - \tilde \vu| }
\]
\[
\leq \delta \| \vu - \tilde \vu \|^2_{L^6(\Omega;R^3)} + c(\delta)
\left\| \left[ \vr \left( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \right) \right]_{\rm res} \right\|_{L^{6/5}(\Omega)}^2
\]
for any $\delta > 0$, where, furthermore,
\[
\left|
\left[ \vr \left( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \right) \right]_{\rm res}
\right| \leq c \left[ \vr + \vr s(\vr, \vt) \right]_{\rm res} .
\]
In accordance with (\ref{m6}), (\ref{m7}),
\bFormula{4.5--} 0 \leq \vr s(\vr, \vt) \leq c \left( \vt^3 + \vr
S \left( \frac{\vr}{\vt^{3/2}} \right) \right) , \eF
where
{
\bFormula{4.5-} \vr S \left( \frac{\vr}{\vt^{3/2}} \right) \leq
c(\vr + \vr [\log\vt]^+ + \vr |\log(\vr)| ). \eF
}
On the other hand, as a
direct consequence of hypotheses (\ref{m3} - \ref{m5}), we get
\bFormula{w5} \vr e(\vr, \vt) \geq c( \vr^{5/3} + \vt^4 ). \eF
{Finally, we observe that (\ref{m13}-\ref{2.15}) imply
\bFormula{4.5+}
t \mapsto \int_\Omega {\cal E}(\vr,\vt|\tilde\vr,\tilde\vt) \ \dx \in
L^\infty(0,T).
\eF
Consequently, using (\ref{w1}), estimates (\ref{4.5--}-\ref{w5}) and the Holder inequality,
we may
infer that
}
\bFormula{w6} \left\| \left[ \vr \left( s(\vr, \vt) -
s(\tilde \vr, \tilde \vt) \right) \right]_{\rm res}
\right\|_{L^{6/5}(\Omega)}^2 \leq c \left( \intO{ \mathcal{E}(\vr,
\vt | \tilde \vr, \tilde \vt ) } \right)^{5/3}. \eF
Combining (\ref{w4}), (\ref{w6})
{
with (\ref{4.5+}) we conclude that \bFormula{w7} \left| \intO{ \vr
\Big( s(\vr, \vt) - s(\tilde \vr, \tilde \vt) \Big) \Big( \tilde
\vu - \vu \Big) \cdot \Grad \tilde \vt } \right| \eF
\[
\leq \| \Grad \tilde \vt \|_{L^\infty(\Omega;R^3)} \left[ \delta
\| \vu - \tilde \vu \|^2_{W^{1,2}_0 (\Omega;R^3)} + c(\delta)
\intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde \vt ) }
\right]\le
\]
\[
\delta \| \vu - \tilde \vu \|^2_{W^{1,2}_0 (\Omega;R^3)} +
K(\delta,\cdot) \intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde
\vt ) }
\]
for any $\delta > 0$. {Here and hereafter, $K(\delta,\cdot)$
is a generic constant depending on $\delta$, $\tilde\vr$, $\tilde\vc u$,
$\tilde\vt$ through the norms induced by (\ref{2.15}),
(\ref{2.16}) and $\underline\vr$, $\underline\vt$, while $K(\cdot)$
is independent of $\delta$ but depends on $\tilde\vr$, $\tilde\vc
u$, $\tilde\vt$, $\underline\vr$, $\underline\vt$ through the
norms (\ref{2.15}), (\ref{2.16}).}
}
\medskip
{\bf Step 3:}
Writing
\[
\intO{ \vr \Big( \partial_t \tilde \vu + \tilde \vu \cdot \Grad \tilde \vu \Big)
\cdot (\tilde \vu - \vu) } = \intO{ \frac{\vr}{\tilde \vr} (\tilde \vu - \vu) \cdot \left(
\Div \tn{S}(\tilde \vt, \Grad \tilde \vu) - \Grad p(\tilde \vr, \tilde \vt) \right) }
\]
\[
= \intO{ \frac{1}{\tilde \vr} (\vr - \tilde \vr) (\tilde \vu - \vu) \cdot \left(
\Div \tn{S}(\tilde \vt, \Grad \tilde \vu) - \Grad p(\tilde \vr, \tilde \vt) \right) }
+ \intO{ (\tilde \vu - \vu) \cdot \left(
\Div \tn{S}(\tilde \vt, \Grad \tilde \vu) - \Grad p(\tilde \vr, \tilde \vt) \right) },
\]
we observe that the first integral on the right-hand side can be
handled in the same way as in Step 2,
{
namely,
$$
\intO{ \left[\frac{1}{\tilde \vr} (\vr - \tilde \vr) (\tilde \vu - \vu) \cdot \left(
\Div \tn{S}(\tilde \vt, \Grad \tilde \vu) - \Grad p(\tilde \vr,
\tilde \vt) \right)\right]_{\rm ess} }\le
K(\delta, \cdot)\left\|\left[\vr-\tilde\vr\right]_{\rm
ess}\right\|^2_{L^2(\Omega)}+\delta\|\vc u-\tilde\vc
u\|^2_{L^2(\Omega;R^3)},
$$
$$
\intO{ \left[\frac{1}{\tilde \vr} (\vr - \tilde \vr) (\tilde \vu - \vu) \cdot \left(
\Div \tn{S}(\tilde \vt, \Grad \tilde \vu) - \Grad p(\tilde \vr,
\tilde \vt) \right)\right]_{\rm res }}
$$
$$
\le
K(\delta, \cdot)\left(\left\|\left[\vr\right]_{\rm
res}\right\|_{L^{6/5}(\Omega)}^2 + \left\|\left[1\right]_{\rm
res}\right\|^2_{L^{6/5}(\Omega)}\right)+\delta\|\vc u-\tilde\vc
u\|^2_{L^6(\Omega;R^3)},
$$
while, integrating by parts,
\[
\intO{ (\tilde \vu - \vu) \cdot \left(
\Div \tn{S}(\tilde \vt, \Grad \tilde \vu) - \Grad p(\tilde \vr, \tilde \vt) \right) }
\]
\[
= \intO{ \left(
\tn{S}(\tilde \vt, \Grad \tilde \vu) : \Grad (\vu - \tilde \vu) + p(\tilde \vr, \tilde \vt)
\Div (\tilde \vu - \vu ) \right) }
\]
Thus using again (\ref{w1}), (\ref{4.5+}) and continuous imbedding
$W^{1,2}(\Omega)\hookrightarrow L^6(\Omega)$ we arrive at
}
\bFormula{w8} \left| \intO{ \vr \Big(
\partial_t \tilde \vu + \tilde \vu \cdot \Grad \tilde \vu \Big)
\cdot (\tilde \vu - \vu) } \right| \leq \intO{ \left(
\tn{S}(\tilde \vt, \Grad \tilde \vu) : \Grad (\vu - \tilde \vu) +
p(\tilde \vr, \tilde \vt) \Div (\tilde \vu - \vu ) \right) } \eF
\[
+ c\left( |\Grad \tilde \vr|, |\Grad \tilde \vt |, |\nabla^2_x
\tilde \vu | \right) \left[ \delta \| \vu - \tilde \vu
\|^2_{W^{1,2}_0(\Omega;R^3)} + c(\delta) \intO{ \mathcal{E}(\vr,
\vt | \tilde \vr, \tilde \vt ) } \right]\le
\]
{
\[
\intO{ \left( \tn{S}(\tilde \vt, \Grad \tilde \vu) : \Grad (\vu -
\tilde \vu) + p(\tilde \vr, \tilde \vt) \Div (\tilde \vu - \vu )
\right) }
\]
\[
+ \delta \| \vu - \tilde \vu \|^2_{W^{1,2}(\Omega;R^3)}
+ K(\delta,\cdot) \intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde
\vt ) }
\]
}
for any $\delta > 0$.
\medskip
{\bf Step 4:}
Next, we get
\[
\intO{ \vr \Big( s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big) \partial_t \tilde \vt } =
\]
\[
\intO{ \vr \Big[ s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big]_{\rm ess} \partial_t \tilde \vt } + \intO{ \vr \Big[ s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big]_{\rm res} \partial_t \tilde \vt },
\]
where
{
\bFormula{w9} \left| \intO{ \vr \Big[ s(\vr, \vt ) -
s(\tilde \vr, \tilde \vt) \Big]_{\rm res} \partial_t \tilde \vt }
\right|
\eF
\[
\leq \| \partial_t \tilde \vt
\|_{L^\infty(\Omega)}
\Big(\int_\Omega\left[\vr s(\vr,\vt)\right]_{\rm res}{\rm d}x
+\|s(\tilde\vr,\tilde\vt)\|_{L^\infty(\Omega)}\int_\Omega[\vr]_{\rm
res}{\rm d}x\Big)\le
K(\cdot) \intO{ \mathcal{E}(\vr, \vt| \tilde
\vr, \tilde \vt ) }, \]
}
while
\[
\intO{ \vr \Big[ s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big]_{\rm ess} \partial_t \tilde \vt }
\]
\[
= \intO{ ( \vr - \tilde \vr) \Big[ s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big]_{\rm ess} \partial_t \tilde \vt } + \intO{ \tilde \vr \Big[ s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big]_{\rm ess} \partial_t \tilde \vt },
\]
where,
{
with help of Taylor-Lagrange formula,
\bFormula{w10} \left| \intO{ ( \vr - \tilde \vr) \Big[ s(\vr, \vt
) - s(\tilde \vr, \tilde \vt) \Big]_{\rm ess}
\partial_t \tilde \vt } \right|
\eF
\[
\leq
\Big(\sup_{(\vr,\vt)\in
[\underline\vr,\overline\vr]\times[\underline\vt,\overline\vt]}
|\partial_\vr s(\vr,\vt)|+ \sup_{(\vr,\vt)\in
[\underline\vr,\overline\vr]\times[\underline\vt,\overline\vt]}|\partial_\vt
s(\vr,\vt)|\Big)\;\|
\partial_t \tilde \vt
\|_{L^\infty(\Omega)}\times \]
\[
\times \int_\Omega\Big[\Big|[\vr-\tilde\vr]_{\rm
ess}\Big|\Big(\Big|[\vr-\tilde\vr]_{\rm ess}\Big|+
\Big|[\vt-\tilde\vt]_{\rm ess}\Big| \Big)\Big]{\rm d}x \le
K(\cdot)\intO{ \mathcal{E}(\vr, \vt| \tilde \vr, \tilde \vt ) }.
\]
}
Finally, we write
\[
\intO{ \tilde \vr \Big[ s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big]_{\rm ess} \partial_t \tilde \vt } = \intO{ \tilde \vr \Big[ s(\vr, \vt) - \partial_\vr s(\tilde \vr, \tilde \vt)
(\vr - \tilde \vr) - \partial_\vt s(\tilde \vr, \tilde \vt)
(\vt - \tilde \vt) - s( \tilde \vr, \tilde \vt) \Big]_{\rm ess} \partial_t \tilde \vt }
\]
\[
- \intO{ \tilde \vr \Big[ \partial_\vr s(\tilde \vr, \tilde \vt)
(\vr - \tilde \vr) + \partial_\vt s(\tilde \vr, \tilde \vt)
(\vt - \tilde \vt) \Big]_{\rm res} \partial_t \tilde \vt } + \intO{ \tilde \vr \Big[ \partial_\vr s(\tilde \vr, \tilde \vt)
(\vr - \tilde \vr) + \partial_\vt s(\tilde \vr, \tilde \vt)
(\vt - \tilde \vt) \Big] \partial_t \tilde \vt },
\]
where the first two integrals on the right-hand side can be
estimated exactly as in (\ref{w9}), (\ref{w10}).
{Thus we conclude that \bFormula{w11} - \intO{ \vr \Big( s(\vr, \vt
) - s(\tilde \vr, \tilde \vt) \Big) \partial_t \tilde \vt } \leq
K(\cdot) \intO{ \mathcal{E}(\vr, \vt| \tilde \vr, \tilde \vt ) }
\eF
}
\[
- \intO{ \tilde \vr \Big[ \partial_\vr s(\tilde \vr, \tilde \vt)
(\vr - \tilde \vr) + \partial_\vt s(\tilde \vr, \tilde \vt)
(\vt - \tilde \vt) \Big] \partial_t \tilde \vt }.
\]
\medskip
{\bf Step 5:}
Similarly to Step 4, we get \bFormula{w12} - \intO{ \vr \Big(
s(\vr, \vt ) - s(\tilde \vr, \tilde \vt) \Big) \tilde \vu \cdot
\Grad \tilde \vt } \leq K(\cdot) \intO{ \mathcal{E}(\vr, \vt|
\tilde \vr, \tilde \vt ) } \eF
\[
- \intO{ \tilde \vr \Big[ \partial_\vr s(\tilde \vr, \tilde \vt)
(\vr - \tilde \vr) + \partial_\vt s(\tilde \vr, \tilde \vt) (\vt -
\tilde \vt) \Big] \tilde \vu \cdot \Grad \tilde \vt }.
\]
\medskip
{\bf Step 6:}
Finally, we have
\bFormula{w13}
\intO{ \left( \left( 1 - \frac{\vr}{\tilde \vr} \right) \partial_t p(\tilde \vr, \tilde \vt)
- \frac{\vr}{\tilde \vr} \vu \cdot \Grad p(\tilde \vr, \tilde \vt) \right) }
\eF
\[
= \intO{ \left(\tilde \vr - \vr \right) \frac{1}{\tilde \vr} \left(
\partial_t p(\tilde \vr, \tilde \vt) + \tilde \vu \cdot \Grad p(\tilde \vr, \tilde \vt) \right) } + \intO{ p(\tilde \vr, \tilde \vt) \Div \vu }
\]
\[
+ \intO{ (\vr - \tilde \vr) \frac{1}{\tilde \vr} \Grad p(\tilde \vr, \tilde \vt)\cdot
(\vu - \tilde \vu) },
\]
where, by means of the same arguments as in Step 2,
\[ \left|
\intO{ (\vr - \tilde \vr) \frac{1}{\tilde \vr} \Grad p(\tilde \vr,
\tilde \vt)\cdot (\vu - \tilde \vu) } \right|\le \]
\[
c\left( |\Grad \tilde \vr|, |\Grad \tilde \vt| \right) \left[\delta \| \vu - \tilde \vu \|^2_{W^{1,2}(\Omega;R^3)} +
\intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde \vt ) } \right]
\]
for any $\delta > 0$.
{
Resuming this step, we have \bFormula{w14} \intO{ \left( \left( 1
- \frac{\vr}{\tilde \vr} \right) \partial_t p(\tilde \vr, \tilde
\vt) - \frac{\vr}{\tilde \vr} \vu \cdot \Grad p(\tilde \vr, \tilde
\vt) \right) } \le \eF
\[
\intO{ \left(\tilde \vr - \vr \right) \frac{1}{\tilde \vr}
\left(
\partial_t p(\tilde \vr, \tilde \vt) + \tilde \vu \cdot \Grad p(\tilde \vr, \tilde \vt) \right) }
+ \intO{ p(\tilde \vr, \tilde \vt) \Div \vu } +
\]
\[
\delta \| \vu - \tilde \vu \|^2_{W^{1,2}(\Omega;R^3)} +
K(\delta,\cdot) \intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde
\vt ) } .
\]
}
\medskip
{\bf Step 7:}
Summing up the estimates (\ref{w3}), (\ref{w7}), (\ref{w8}), (\ref{w11} - \ref{w14}),
we can rewrite the relative entropy inequality (\ref{w2}) in the form
\bFormula{w15}
\intO{ \left( \frac{1}{2} \vr |\vu - \tilde \vu |^2
+ \mathcal{E}(\vr, \vt| \tilde \vr, \tilde \vt ) \right) (\tau, \cdot) }
\eF
\[
+ \int_0^\tau \intO{ \left( \frac{\tilde \vt}{\vt} \tn{S} (\vt, \Grad \vu) : \Grad \vu -
\tn{S} (\tilde \vt, \Grad \tilde \vu ): (\Grad \vu - \Grad \tilde \vu ) -
\tn{S}(\vt, \Grad \vu) : \Grad \tilde \vu \right) } \ \dt
\]
\[
+ \int_0^\tau \intO{ \left( \frac{\vc{q} (\vt, \Grad \vt) \cdot
\Grad \tilde \vt }{\vt} - \frac{\tilde \vt}{\vt} \frac{\vc{q}
(\vt, \Grad \vt) \cdot \Grad \vt }{\vt} \right) } \ \dt\leq
\]
{
\[
\int_0^\tau \left[ \delta \| \vu - \tilde \vu
\|^2_{W^{1,2}(\Omega;R^3)} + K(\delta,\cdot)\intO{ \left(
\frac{1}{2} \vr | \vu - \tilde \vu |^2 + \mathcal{E}(\vr, \vt |
\tilde \vr, \tilde \vt ) \right) } \right] \ \dt
\]
}
\[
+ \int_0^\tau \intO{ \Big( p(\tilde \vr, \tilde \vt) - p(\vr, \vt) \Big) \Div \tilde \vu }
\ \dt
\]
\[
+ \int_0^\tau \intO{ (\tilde \vr - \vr) \frac{1}{\tilde \vr} \left[ \partial_t p(\tilde \vr, \tilde \vt) + \tilde \vu \cdot \Grad p(\tilde \vr, \tilde \vt) \right] } \ \dt
\]
\[
- \int_0^\tau \intO{ \tilde \vr \left( \partial_\vr s(\tilde \vr, \tilde \vt) (\vr - \tilde \vr) + \partial_\vt s(\tilde \vr, \tilde \vt) (\vt - \tilde \vt) \right) \left[
\partial_t \tilde \vt + \tilde \vu \cdot \Grad \tilde \vt \right] } \ \dt
\]
for any $\delta > 0$.
\medskip
{\bf Step 8:}
Our next goal is to control the last three integrals on the right-hand side of (\ref{w15}).
To this end, we use (\ref{sim}) to obtain
\[
\intO{ (\tilde \vr - \vr) \frac{1}{\tilde \vr} \left[ \partial_t p(\tilde \vr, \tilde \vt) + \tilde \vu \cdot \Grad p(\tilde \vr, \tilde \vt) \right] }
\]
\[
- \intO{ \tilde \vr \left( \partial_\vr s(\tilde \vr, \tilde \vt) (\vr - \tilde \vr) + \partial_\vt s(\tilde \vr, \tilde \vt) (\vt - \tilde \vt) \right) \left[
\partial_t \tilde \vt + \tilde \vu \cdot \Grad \tilde \vt \right] }
\]
\[
= \intO{ \tilde \vr (\tilde \vt - \vt) \partial_\vt s(\tilde \vr, \tilde \vt)
\left[ \partial_t \tilde \vt + \tilde \vu \cdot \Grad \tilde \vt \right] } +
\intO{ (\tilde \vr - \vr ) \frac{1}{\tilde \vr} \partial_\vr p(\tilde \vr, \tilde \vt)
\left[ \partial_t \tilde \vr + \tilde \vu \cdot \Grad \tilde \vr \right] },
\]
where, as $\tilde \vr$, $\tilde \vu$ satisfy the equation of continuity (\ref{i1}),
\bFormula{w16}
\intO{ (\tilde \vr - \vr ) \frac{1}{\tilde \vr} \partial_\vr p(\tilde \vr, \tilde \vt)
\left[ \partial_t \tilde \vr + \tilde \vu \cdot \Grad \tilde \vr \right] } = -
\intO{ (\tilde \vr - \vr) \partial_\vr p(\tilde \vr, \tilde \vt) \Div \tilde \vu }.
\eF
Finally, using (\ref{sim}) once more, we deduce that
\bFormula{w17}
\intO{ \tilde \vr (\tilde \vt - \vt) \partial_\vt s(\tilde \vr, \tilde \vt)
\left[ \partial_t \tilde \vt + \tilde \vu \cdot \Grad \tilde \vt \right] }
\eF
\[
\intO{ \tilde \vr (\tilde \vt - \vt)
\left[ \partial_t s(\tilde \vr, \tilde \vt) + \tilde \vu \cdot \Grad s(\tilde \vr, \tilde \vt) \right] } - \intO{ (\tilde \vt - \vt)
\partial_\vt p(\tilde \vr, \tilde \vt) \Div \tilde \vu }
\]
\[
= \intO{ (\tilde \vt - \vt) \left[ \frac{1}{\tilde \vt} \left(
\tn{S} (\tilde \vt, \Grad \tilde \vu) : \Grad \tilde \vu - \frac{\vc{q}(\tilde \vt, \Grad \tilde \vt) \cdot \Grad \tilde \vt}{\tilde \vt} \right) - \Div
\left( \frac{\vc{q} (\tilde \vt, \Grad \tilde \vt)}{\tilde \vt} \right) \right] }
\]
\[
- \intO{ (\tilde \vt - \vt)
\partial_\vt p(\tilde \vr, \tilde \vt) \Div \tilde \vu }
\]
{Seeing that
\[
\left| \intO{ \left( p(\tilde \vr, \tilde \vt) - \partial_\vr
p(\tilde \vr, \tilde \vt) (\tilde \vr - \vr) - \partial_\vt
p(\tilde \vr, \tilde \vt) (\tilde \vt - \vt) - p(\vr, \vt) \right)
\Div \tilde\vu } \right|
\]
}
\[
\leq
c \| \Div \tilde \vu \|_{L^\infty (\Omega)} \intO{ \mathcal{E} (\vr, \vt | \tilde \vr, \tilde \vt ) }
\]
we may use the previous relations to rewrite (\ref{w15}) in the form
\bFormula{w18}
\intO{ \left( \frac{1}{2} \vr |\vu - \tilde \vu |^2 + \mathcal{E}(\vr, \vt| \tilde \vr, \tilde \vt ) \right) (\tau, \cdot) }
\eF
\[
+ \int_0^\tau \intO{ \left( \frac{\tilde \vt}{\vt} \tn{S} (\vt, \Grad \vu) : \Grad \vu -
\tn{S} (\tilde \vt, \Grad \tilde \vu ): (\Grad \vu - \Grad \tilde \vu ) -
\tn{S}(\vt, \Grad \vu) : \Grad \tilde \vu - \frac{\tilde \vt - \vt}{\tilde \vt}
\tn{S}(\tilde \vt, \Grad \tilde \vu ) : \Grad \tilde \vu \right) } \ \dt
\]
\[
+ \int_0^\tau \intO{ \left( \frac{\vc{q} (\vt, \Grad \vt) \cdot \Grad \tilde \vt }{\vt} -
\frac{\tilde \vt}{\vt} \frac{\vc{q} (\vt, \Grad \vt) \cdot \Grad \vt }{\vt}
+ (\tilde \vt - \vt) \frac{ \vc{q}(\tilde \vt, \Grad \tilde \vt) \cdot \Grad \tilde \vt }{
{\tilde \vt}^2 } + \frac{\vc{q}(\tilde \vt, \Grad \tilde \vt)}{\tilde \vt} \cdot
\Grad (\vt - \tilde \vt)
\right) }
\ \dt
\]
{
\[
\leq \int_0^\tau \left[ \delta \| \vu - \tilde \vu
\|^2_{W^{1,2}(\Omega;R^3)} + K(\delta,\cdot) \intO{ \left(
\frac{1}{2} \vr | \vu - \tilde \vu |^2 + \mathcal{E}(\vr, \vt |
\tilde \vr, \tilde \vt ) \right) } \right] \ \dt
\]
}
for any $\delta > 0$.
\subsection{Dissipative terms}
Our ultimate goal in the proof of Theorem \ref{Tm1} is to show that the ``dissipative'' terms appearing on the left-hand side of (\ref{w18}) containing $\Grad \vu$, $\Grad \vt$ are strong enough to control the $W^{1,2}-$norm of the velocity.
\subsubsection{Viscosity}
In accordance with hypothesis (\ref{m8}), we have
\[
\tn{S}(\vt, \Grad \vu) = \tn{S}^0 (\vt, \Grad \vu) + \tn{S}^1(\vt, \Grad \vu),
\]
where
\[
\tn{S}^0 (\vt, \Grad \vu) = \mu_0 \Big(\Grad \vu + \Grad^t \vu - \frac{2}{3} \Div \vu \tn{I} \Big), \ \tn{S}^1 (\vt, \Grad \vu) = \mu_1 \vt \Big(\Grad \vu + \Grad^t \vu - \frac{2}{3} \Div \vu \tn{I} \Big).
\]
Now, we write
\[
\frac{\tilde \vt}{\vt} \tn{S}^1 (\vt, \Grad \vu) : \Grad \vu -
\tn{S}^1 (\tilde \vt, \Grad \tilde \vu ): (\Grad \vu - \Grad \tilde \vu ) -
\tn{S}^1(\vt, \Grad \vu) : \Grad \tilde \vu - \frac{\tilde \vt - \vt}{\tilde \vt}
\tn{S}^1(\tilde \vt, \Grad \tilde \vu ) : \Grad \tilde \vu
\]
\[
= \tilde \vt \left( \frac{\tn{S}^1 (\vt, \Grad \vu) }{\vt} - \frac{\tn{S}^1 (\tilde \vt,
\Grad \tilde \vu) }{\tilde \vt} \right) : (\Grad \vu - \Grad \tilde \vu )
+ (\tilde \vt - \vt ) \left( \frac{\tn{S}^1 (\vt, \Grad \vu) }{\vt} - \frac{\tn{S}^1 (\tilde \vt,
\Grad \tilde \vu) }{\tilde \vt} \right) : \Grad \tilde \vu,
\]
where, by virtue of Korn's inequality, \bFormula{w19} \intO{
\tilde \vt \left( \frac{\tn{S}^1 (\vt, \Grad \vu) }{\vt} -
\frac{\tn{S}^1 (\tilde \vt, \Grad \tilde \vu) }{\tilde \vt}
\right) : (\Grad \vu - \Grad \tilde \vu ) } \geq c \mu_1 \| \vu -
\tilde \vu \|^2_{W^{1,2}(\Omega;R^3)}. \eF
On the other hand, similarly to the preceding part, we can show that
\bFormula{w20}
\left| \intO{ (\tilde \vt - \vt ) \left( \frac{\tn{S}^1 (\vt, \Grad \vu) }{\vt} - \frac{\tn{S}^1 (\tilde \vt,
\Grad \tilde \vu) }{\tilde \vt} \right) : \Grad \tilde \vu } \right|
\eF
\[
\leq \| \Grad \tilde \vu \|_{L^\infty(\Omega; R^{3 \times 3})}
\left( \delta \| \vu - \tilde \vu \|^2_{W^{1,2} (\Omega;R^3)} +
c(\delta) \Big(\left\| [\vt-\tilde\vt]_{\rm
ess}\right\|^2_{L^2(\Omega)}+ \left\| [\vt-\tilde\vt]_{\rm
res}\right\|^2_{L^2(\Omega)}\Big)\right)
\]
\[
\leq \delta \| \vu - \tilde \vu \|^2_{W^{1,2} (\Omega;R^3)} +
K(\delta,\cdot) \intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde
\vt ) }
\]
for any $\delta > 0$, where we have used again (\ref{w1}).
{Next,
\[
\frac{\tilde \vt}{\vt} \tn{S}^0 (\Grad \vu) : \Grad \vu -
\tn{S}^0 (\Grad \tilde \vu ): (\Grad \vu - \Grad \tilde \vu ) -
\tn{S}^0(\Grad \vu) : \Grad \tilde \vu - \frac{\tilde \vt - \vt}{\tilde \vt}
\tn{S}^0(\Grad \tilde \vu ) : \Grad \tilde \vu
\]
\[
= \frac{\tilde \vt}{\vt} \Big( \tn S^0 (\Grad \vu) - \tn S^0
(\Grad \tilde \vu) \Big) : \Grad (\vu - \tilde \vu ) + \tilde \vt
\left( \frac{1}{\vt} - \frac{1}{\tilde \vt} \right) \tn{S}^0
(\Grad \tilde \vu) : \Grad (\vu - \tilde \vu).
\]
First suppose that $\vt \geq \tilde \vt$. Since the function $\vt
\mapsto 1/\vt$ is Lipschitz on the set $\vt \geq \tilde \vt$, we
conclude that
\bFormula{w21} \int_{ \{ \vt \geq \tilde \vt \} } \left| \tilde
\vt \left( \frac{1}{\vt} - \frac{1}{\tilde \vt} \right) \tn{S}^0
(\Grad \tilde \vu) : \Grad (\vu - \tilde \vu) \right| + \left|
\frac{ \tilde \vt - \vt }{\vt} \Big( \tn S^0(\Grad \vu) - \tn
S^0(\Grad \tilde \vu) \Big) : \Grad \tilde \vu \right| \ \dx \eF
\[
\leq \Big(\frac 1{\underline
\vt}+\|\Grad\tilde\vu\|_{L^\infty(\Omega;\R^3)}+\|\tilde\vt\|_{L^\infty(\Omega)}\Big)
\Big( \delta \| \Grad \vu - \Grad \tilde \vu \|^2_{L^2(\Omega;R^{3
\times 3})}+c(\delta)\|\vt-\tilde\vt\|^2_{L^2(\Omega)}\Big)
\]
\[
\delta \| \Grad \vu - \Grad \tilde \vu \|^2_{L^2(\Omega;R^{3
\times 3})} + K(\delta,\cdot) \intO{ \mathcal{E}(\vr, \vt | \tilde
\vr, \tilde \vt ) }
\]
for any $\delta > 0$.
}
Finally, if $0 < \vt \leq \tilde \vt$, we have
\[
\frac{\tilde \vt}{\vt} \tn{S}^0 (\Grad \vu) : \Grad \vu -
\tn{S}^0 (\Grad \tilde \vu ): (\Grad \vu - \Grad \tilde \vu ) -
\tn{S}^0(\Grad \vu) : \Grad \tilde \vu - \frac{\tilde \vt - \vt}{\tilde \vt}
\tn{S}^0(\Grad \tilde \vu ) : \Grad \tilde \vu
\]
\[
\geq \left( \tn{S}^0 (\Grad \vu) - \tn{S}^0 (\Grad \tilde \vu) \right): (\Grad \vu - \Grad \tilde \vu) + \frac{\tilde \vt - \vt}{\tilde \vt} \left[ \tn{S}^0 (\Grad \vu) : \Grad \vu -
\tn{S}^0 (\Grad \tilde \vu) : \Grad \tilde \vu \right];
\]
whence, by means of convexity of the function $\Grad \vu \mapsto \tn{S}^0 (\Grad \vu) : \Grad \vu$,
\[
\frac{\tilde \vt - \vt}{\tilde \vt} \left[ \tn{S}^0 (\Grad \vu) :
\Grad \vu - \tn{S}^0 (\Grad \tilde \vu) : \Grad \tilde \vu \right]
\geq \frac{\tilde \vt - \vt}{\tilde \vt} \tn{S}^0 (\Grad \tilde
\vu) \cdot \Grad (\vu - \tilde \vu),
\]
{
where, similarly as in (\ref{w21}),
\[
\int_\Omega\frac{\tilde \vt - \vt}{\tilde \vt} \tn{S}^0 (\Grad
\tilde \vu) \cdot \Grad (\vu - \tilde \vu){\rm d}x \le
K(\delta,\cdot) \intO{ \mathcal{E}(\vr, \vt | \tilde \vr, \tilde
\vt )}.
\]
}
Summing up the results of the section, we may choose $\delta > 0$ so small that relation
(\ref{w18}) takes the form
\bFormula{w22}
\intO{ \left( \frac{1}{2} \vr |\vu - \tilde \vu |^2 +
\mathcal{E}(\vr, \vt| \tilde \vr, \tilde \vt ) \right) (\tau, \cdot) } + c_1 \int_0^\tau \intO{ | \Grad \vu - \Grad \tilde \vu |^2 } \ \dt
\eF
\[
+ \int_0^\tau \intO{ \left( \frac{\vc{q} (\vt, \Grad \vt) \cdot \Grad \tilde \vt }{\vt} -
\frac{\tilde \vt}{\vt} \frac{\vc{q} (\vt, \Grad \vt) \cdot \Grad \vt }{\vt}
+ (\tilde \vt - \vt) \frac{ \vc{q}(\tilde \vt, \Grad \tilde \vt) \cdot \Grad \tilde \vt }{
{\tilde \vt}^2 } + \frac{\vc{q}(\tilde \vt, \Grad \tilde \vt)}{\tilde \vt} \cdot
\Grad (\vt - \tilde \vt)
\right) }
\ \dt
\]
{
\[
\leq c_2 \int_0^\tau \intO{ \left( \frac{1}{2} \vr | \vu - \tilde
\vu |^2 + \mathcal{E}(\vr, \vt | \tilde \vr, \tilde \vt ) \right)
} \ \dt.
\]
}
\subsubsection{Heat conductivity}
In accordance with hypothesis (\ref{m9}), we write
\[
\vc{q} (\vt, \Grad \vt) = - \kappa_0 \Grad \vt - \kappa_2 \vt^2 \Grad \vt - \kappa_3
\vt^3 \Grad \vt.
\]
We compute
\bFormula{w23}
\frac{\tilde \vt}{\vt} \frac{\kappa_0}{\vt} |\Grad \vt |^2 - \frac{\kappa_0}{\vt}
\Grad \vt \cdot \Grad \tilde \vt + \frac{\vt - \tilde \vt}{\tilde \vt} \frac{\kappa_0}{\tilde \vt}
|\Grad \tilde \vt |^2 + \frac{\kappa_0}{\tilde \vt} \Grad \tilde \vt \cdot \Grad (\tilde \vt - \vt)
\eF
\[
= \kappa_0 \Big[ \tilde \vt |\Grad \log(\vt) |^2 - \tilde \vt \Grad \log (\vt) \cdot \Grad \log(\tilde \vt) + (\vt - \tilde \vt) |\Grad \log (\tilde \vt) |^2 + \Grad \log (\tilde \vt) \cdot \Grad (\tilde \vt - \vt) \Big]
\]
\[
= \kappa_0 \Big[ \tilde \vt | \Grad \log(\vt) - \Grad \log (\tilde \vt) |^2 + (\vt - \tilde \vt) |\Grad \log (\tilde \vt) |^2 + \Grad \log (\tilde \vt) \cdot \Grad (\tilde \vt - \vt)
\]
\[
+ \tilde \vt \Grad \log (\tilde \vt) \cdot \Big( \Grad \log(\vt) - \Grad \log(\tilde \vt) \Big) \Big] = \kappa_0 \Big[ \tilde \vt | \Grad \log(\vt) - \Grad \log (\tilde \vt) |^2
\]
\[
+ (\vt - \tilde \vt) |\Grad \log (\tilde \vt) |^2 + (\tilde \vt - \vt) \Grad \log(\tilde \vt) \cdot \Grad \log(\vt) \Big]
\]
\[
= \kappa_0 \left[ \tilde \vt | \Grad \log(\vt) - \Grad \log (\tilde \vt) |^2 + { (\vt - \tilde \vt) \Grad \log(\tilde \vt) \cdot \Grad \Big( \log(\tilde \vt ) - \log(\vt) \Big)} \right],
\]
where the second term on the right-hand side can be ``absorbed'' by the remaining integrals in (\ref{w22}).
Similarly, we get
\bFormula{w24}
\kappa_2 \tilde \vt |\Grad \vt |^2 - \kappa_2 \vt
\Grad \vt \cdot \Grad \tilde \vt + \kappa_2 (\vt - \tilde \vt)
|\Grad \tilde \vt |^2 + \kappa_2 \tilde \vt \Grad \tilde \vt \cdot \Grad (\tilde \vt - \vt)
\eF
\[
= \kappa_2 \Big[ \tilde \vt | \Grad \vt - \Grad \tilde \vt |^2 + {(\vt - \tilde \vt) \Grad \tilde \vt \cdot \Grad (\vt - \tilde \vt)} \Big].
\]
Finally,
\bFormula{w25}
\kappa_3 \vt \tilde \vt | \Grad \vt |^2 - \kappa_3 \vt^2
\Grad \vt \cdot \Grad \tilde \vt + \kappa_3 (\vt - \tilde \vt) \tilde \vt
|\Grad \tilde \vt |^2 + \kappa_3 {\tilde \vt}^2 \Grad \tilde \vt \cdot \Grad (\tilde \vt - \vt)
\eF
\[
= \kappa_3 \Big[ \tilde \vt \vt \Grad \vt \cdot (\Grad \vt - \Grad \tilde \vt ) +
\tilde \vt \vt \Grad \vt \cdot \Grad \tilde \vt - \vt^2 \Grad \vt \cdot \Grad \tilde \vt +
(\vt - \tilde \vt) \tilde \vt | \Grad \tilde \vt |^2 + {\tilde \vt}^2 \Grad \tilde \vt \cdot \Grad (\tilde \vt - \vt) \Big]
\]
\[
= \kappa_3 \Big[ \tilde \vt \vt | \Grad \vt - \Grad \tilde \vt |^2 + 2 \vt \tilde \vt \Grad \vt \cdot \Grad \tilde \vt - \vt^2 \Grad \vt \cdot \Grad \tilde \vt - \tilde \vt^2 \Grad \vt
\cdot \Grad \tilde \vt \Big]
\]
\[
= \kappa_3 \tilde \vt \vt | \Grad \vt - \Grad \tilde \vt |^2 - \kappa_3 (\vt - \tilde \vt)^2 \Grad \vt \cdot \Grad \tilde \vt,
\]
where
\[
(\vt - \tilde \vt)^2 \Grad \vt \cdot \Grad \tilde \vt = |\Grad \tilde \vt |^2 (\vt - \tilde \vt)^2 + \Grad (\vt - \tilde \vt) \cdot \Grad \tilde \vt (\vt - \tilde \vt)^2.
\]
We conclude by observing that
\[
\Grad (\vt - \tilde \vt) \cdot \Grad \tilde \vt (\vt - \tilde \vt)^2 =
\Grad (\vt - \tilde \vt) \cdot \Grad \tilde \vt [ \vt - \tilde \vt]^2_{\rm ess} +
\Grad (\vt - \tilde \vt) \cdot \Grad \tilde \vt [ \vt - \tilde \vt]^2_{\rm res},
\]
where, furthermore,
\bFormula{w26} \intO{\left| \Grad (\vt - \tilde \vt) \cdot \Grad
\tilde \vt [ \vt - \tilde \vt]^2_{\rm res} \right|} \leq \intO{|
\Grad \tilde \vt | \left[ \delta |\Grad \vt - \Grad \tilde \vt |^2
+ c(\delta) \left[\vt - \tilde \vt \right]^4_{\rm res} \right]}
\eF
{
\[
\le \delta\|\Grad\vu-\Grad\tilde\vu\|^2_{L^2(\Omega;R^{3\times
3})}+ K(\delta,\cdot) \intO{{\cal
E}(\vr,\vt|\tilde\vr,\tilde\vt)},
\]
\[
\intO{\left| \Grad (\vt - \tilde \vt) \cdot \Grad \tilde \vt [ \vt
- \tilde \vt]^2_{\rm ess} \right|} \leq 2\overline\vt\intO{| \Grad
\tilde \vt | \left[ \delta |\Grad \vt - \Grad \tilde \vt | +
c(\delta) \left[ \vt - \tilde \vt \right]^2_{{\rm ess}} \right]}
\]
\[
\le \delta\|\Grad\vu-\Grad\tilde\vu\|^2_{L^2(\Omega;R^{3\times
3})}+ K(\delta,\cdot) \intO{{\cal E}(\vr,\vt|\tilde\vr,\tilde\vt)}
\]
for any $\delta > 0$.
}
Summing up (\ref{w23} - \ref{w26}) we may write (\ref{w22}) as
\bFormula{w27}
\intO{ \left( \frac{1}{2} \vr |\vu - \tilde \vu |^2 +
\mathcal{E}(\vr, \vt| \tilde \vr, \tilde \vt ) \right) (\tau, \cdot) } +
c_1 \int_0^\tau \intO{ | \Grad \vu - \Grad \tilde \vu |^2 } \ \dt
\eF
\[
+ c_2 \left[ \int_0^\tau \intO{ |\Grad \vt - \Grad \tilde \vt |^2
} \ \dt + \int_0^\tau \intO{ | \Grad (\log(\vt)) - \Grad
\log(\tilde \vt) |^2 } \ \dt \right]
\]
\[
\leq c_3 \int_0^\tau \intO{ \left( \frac{1}{2} \vr | \vu - \tilde
\vu |^2 + \mathcal{E}(\vr, \vt | \tilde \vr, \tilde \vt ) \right)
} \ \dt \ \mbox{for a.a} \ \tau \in (0,T),
\]
which yields the desired conclusion
\[
\vr \equiv \tilde \vr, \ \vt \equiv \tilde \vt, \ \vu \equiv \tilde \vu.
\]
We have proved Theorem \ref{Tm1}.
\section{Concluding remarks}
\label{c}
The structural restrictions introduced in Section \ref{m}, and, in particular, the presence of the radiation pressure proportional to $\vt^4$ were motivated by the \emph{existence theory} developed in \cite[Chapter 3]{FENO6}. More refined arguments could be used to show that many of these assumptions could be relaxed in the proof of weak-strong uniqueness.
In particular, the estimates based on the presence of the radiation components of the thermodynamic functions $p$, $e$, and $s$ could be performed by means of the ``dissipative'' terms on the left-hand side of (\ref{w27}) combined with some variant of Poincare's inequality.
Similar result can be obtained for other types of boundary conditions and even on unbounded spatial domains.
Last but not least, we remark that the smoothness assumptions imposed on the classical solution $\tilde \vr$, $\tilde \vt$, $\tilde \vu$ can be relaxed in terms of the integrability properties of the weak solution $\vr$, $\vt$, $\vu$.
Our final remark concerns the weak formulation of the Navier-Stokes-Fourier system introduced in Section \ref{ws}. Note that there are at least two alternative ways how to replace the entropy balance (\ref{i3}), namely by the
total energy balance
\bFormula{C1}
\partial_t \left( \frac{1}{2} \vr |\vu|^2 + \vr e(\vr, \vt) \right) +
\Div \left[ \left( \frac{1}{2} \vr |\vu|^2 + \vr e(\vr, \vt) + p(\vr,\vt) \right) \vu \right] + \Div \vc{q} =
\Div (\tn{S} \vu ),
\eF
or by the internal energy balance
\bFormula{C2}
\partial_t (\vr e(\vr, \vt)) + \Div (\vr e(\vr, \vt) \vu ) + \Div \vc{q} = \tn{S} : \Grad \vu - p \Div \vu.
\eF
Although (\ref{C1}), (\ref{C2}) are equivalent to (\ref{i3}) for classical solutions, this is, in general, not the case
in the framework of weak solutions. As we have seen, it is precisely the entropy balance (\ref{i3}) that gives rise, in combination with (\ref{i8}), the relative entropy inequality (\ref{d8}) yielding the weak-strong uniqueness property.
This fact may be seen as another argument in favor of the weak formulation of the Navier-Stokes-Fourier system based
on (\ref{i3}), (\ref{i8}).
\def\ocirc#1{\ifmmode\setbox0=\hbox{$#1$}\dimen0=\ht0 \advance\dimen0
by1pt\rlap{\hbox to\wd0{\hss\raise\dimen0
\hbox{\hskip.2em$\scriptscriptstyle\circ$}\hss}}#1\else {\accent"17 #1}\fi}
|
\section{Introduction}
Classical time series analysis has generally focused on a single (potentially multivariate) time series from which inferences are to be made. For example, one might monitor the daily returns of a particular stock index and wish to infer the changing regimes of volatility. However, in a growing number of fields, interest is in making inferences based on a \emph{collection} of related time series. One might monitor multiple financial indices, or collect EEG data from a given patient at multiple non-contiguous epochs. We focus on time series with dynamics that are too complex to be described using standard linear dynamical models (e.g., autoregressive processes), but that exhibit switches among a set of \emph{behaviors} that describe locally coherent and simple dynamic modes that persist over a segment of time. For example, stock returns might be modeled via switches between regimes of volatility or an EEG recording between spiking patterns dependent on seizure type. In such cases, one would like to discover and model the dynamical behaviors which are shared among several related time series. In essence, we would like to capture a combinatorial form of shrinkage involving subsets of behaviors from an overall library of behaviors.
As a specific motivating example that we consider later in this paper, consider a multivariate time series that arises when position and velocity sensors are placed on the limbs and joints of a person who is going through an exercise routine. In the specific dataset that we analyze, the time series can be segmented into types of exercise (e.g., jumping jacks, touch-the-toes and twists). The goal is to discover these exercise types (i.e., the ``behaviors'') and their occurrences in the data stream. Moreover, the overall dataset consists of multiple time series obtained from multiple individuals, each of whom performs some subset of exercise types. We would like to take advantage of the overlap between individuals, such that if a ``jumping-jack behavior'' is discovered in the time series for one individual then it can be used in modeling the data for other individuals.
A flexible yet simple method of describing single time series with such patterned behaviors is the class of \emph{Markov switching processes}. These processes assume that the time series can be described via Markov transitions between a set of latent dynamic behaviors which are individually modeled via temporally independent or linear dynamical systems. Examples include the hidden Markov model (HMM), switching vector autoregressive (VAR) process, and switching linear dynamical system (SLDS). These models have proven useful in such diverse fields as speech recognition, econometrics, neuroscience, remote target tracking, and human motion capture. In this paper, we focus our attention on the descriptive yet computationally tractable class of switching VAR processes. In this case, the state, or \emph{dynamical mode}, of the underlying Markov process encodes the dynamic behavior exhibited at a given time step and each dynamic behavior is a VAR process. That is, conditioned on the Markov-evolving state, the likelihood is simply a VAR model.
To discover the dynamic behaviors shared between multiple time series, we propose a feature-based model. Globally, the collection of time series can be described by the shared \emph{library} of possible dynamic behaviors. Individually, however, a given time series will only exhibit some subset of these behaviors. That is, each time series has a \emph{vocabulary} of possible states. The goal in relating the time series is to discover which behaviors are shared amongst the time series and which are unique. Let us represent the vocabulary of time series $i$ by a \emph{feature vector} $\fset{i}$, with $f_{ik}=1$ indicating that time series $i$ has behavior $k$ in its vocabulary. We seek a prior for these feature vectors. We particularly aim to allow flexibility in the number of total and time-series-specific behaviors, and to encourage time series to share similar subsets of the large set of possible behaviors. Our desiderata motivate a feature-based Bayesian nonparametric approach based on the \emph{beta process}~\citep{Hjort:90,Thibaux:07}. Such an approach allows for \emph{infinitely} many potential dynamic behaviors, but encourages a sparse representation.
In our scenario, one can think of the beta process as defining a coin-flipping probability for each of an infinite set of possible dynamic behaviors. Each time series' feature vector is modeled as the result of a Bernoulli process draw: the beta-process-determined coins are flipped for each dynamic behavior and the set of resulting heads indicate the set of selected features (implicitly defining an infinite-dimensional feature vector.) The properties of the beta process induce sparsity in the feature space by encouraging sharing of features among the Bernoulli process observations. Specifically, the total sum of coin weights is finite and only certain dynamic behaviors have large coin weights. Thus, certain dynamic behaviors are more prevalent in the vocabularies of the time series, though the resulting vocabularies clearly need not be identical. As shown by~\citet{Thibaux:07}, integrating over the latent beta process random measure (i.e., coin-flipping weights) induces a predictive distribution on features known as the \emph{Indian buffet process} (IBP)~\citep{GriffithsGhahramani:05}. Computationally, this representation is key. Given a sampled feature set, our model reduces to a collection of finite Bayesian VAR processes with partially shared parameters.
Our presentation is organized as follows. The beta process is reviewed in Section~\ref{background:BetaProcess}, following a brief overview of Markov switching processes. In Section~\ref{sec:model}, we present our proposed beta-process-based model for jointly modeling multiple related Markov switching processes. Efficient posterior computations based on a Markov chain Monte Carlo (MCMC) algorithm are developed in Section~\ref{sec:MCMC}. The algorithm does not rely on model truncation; instead, we exploit the finite dynamical system induced by a fixed set of features to efficiently compute acceptance probabilities, and reversible jump birth and death proposals to explore new features. The sampling of features relies on the IBP interpretation of the beta process---the connection between the beta process and the IBP is outlined in Section~\ref{sec:IBP}. In Section~\ref{sec:related}, we describe related approaches. Section~\ref{sec:synth} examines the benefits of our proposed feature-based model on several synthetic datasets. Finally, in Section~\ref{sec:MoCap} we present promising results on the challenging task of unsupervised segmentation of data from the CMU motion capture database~\citep{CMUmocap}.
\section{Background}
\subsection{Markov Switching Processes}
\label{sec:MarkovSwitchingProcesses}
\subsubsection*{Hidden Markov Models}
The hidden Markov model, or \emph{HMM}, is a class of doubly
stochastic processes based on an underlying, discrete-valued state
sequence that is modeled as Markovian~\citep{Rabiner:89}. Conditioned
on this state sequence, the model assumes that the observations,
which may be discrete or continuous valued, are independent. Specifically,
let $z_t$ denote the state, or \emph{dynamical mode}, of the Markov chain at time~$t$
and let $\pi_j$ denote the state-specific \emph{transition distribution} for
mode $j$. Then, the Markovian structure on the mode sequence
dictates that
\begin{align}
z_t\mid z_{t-1} \sim \pi_{z_{t-1}}. \label{eqn:HMMmode}
\end{align}
Given the mode $z_t$, the observation $y_t$ is a conditionally
independent emission
\begin{align}
y_t \mid z_t \sim F(\theta_{z_t})
\end{align}
for an indexed family of distributions $F(\cdot)$. Here, $\theta_i$
are the \emph{emission parameters} for mode~$i$.
\subsubsection*{Switching VAR Processes}
The modeling assumption of the HMM that observations are conditionally independent given the latent mode sequence is often insufficient in capturing the temporal dependencies present in many datasets. Instead, one can assume that the observations have conditionally \emph{linear} dynamics. The latent HMM dynamical mode then models switches between a set of such linear dynamical systems in order to capture more complex dynamical phenomena. We restrict our attention in this paper to switching vector autoregressive (VAR) processes, or \emph{autoregressive HMMs} (AR-HMMs), which are broadly applicable in many domains while maintaining a number of simplifying properties that make them a practical choice computationally.
We define an AR-HMM, with switches between order-$r$ vector autoregressive processes~\footnote{We denote an order-$r$ VAR process by VAR($r$).}, as
\begin{equation}
\begin{aligned}
\BF{y}_t &= \sum_{i=1}^r A_{i,z_t}\BF{y}_{t-i} + \BF{e}_t(z_t),
\end{aligned}
\label{eqn:SVAR}
\end{equation}
where $z_t$ represents the HMM latent dynamical mode of the system at time $t$, and is defined as in Eq.~\eqref{eqn:HMMmode}. The mode-specific additive noise term is distributed as $\BF{e}_t(z_t) \sim \mathcal{N}(0,\Sigma_{z_t})$. We refer to $\BF{A}_k = \{A_{1,k},\dots,A_{r,k}\}$ as the set of \emph{lag matrices}. Note that the standard HMM with Gaussian emissions arises as a special case of this model when $\BF{A}_{k}=\BF{0}$ for all~$k$.
\subsection{Relating Multiple Time Series}
\label{sec:multipleTimeSeries}
In our applications of interest, we are faced with a \emph{collection} of $N$ time series representing realizations of related dynamical phenomena. We assume that each time series is individually modeled via a switching VAR process, as in Equation~\eqref{eqn:SVAR}. Denote the VAR parameters for the $k^{th}$ dynamical mode as $\theta_k = \{\BF{A}_k,\Sigma_k\}$, and assume that we have an unbounded set of possible VAR models $\{\theta_1,\theta_2,\dots\}$. For example, these parameters might each define a linear motion model for the behaviors \emph{walking}, \emph{running}, \emph{jumping}, and so on; our time series are then each modeled as Markov switches between these behaviors. We will sometimes avail ourselves the convenient shorthand of referring to $k$ itself as a ``behavior,'' where the intended meaning is the VAR model parameterized by $\theta_k$.
The way in which our $N$ time series are related is by the overlap in the set of dynamic behaviors that each exhibits. For example, imagine that our $N$ time series represent observation sequences from the exercise routines of $N$ people. We expect there to be some overlap in the behaviors exhibited, but also some variability---e.g., some people may solely switch between walking and running, while others switch between running and jumping.
One can represent the set of behaviors available to each of the time series models with a list of binary \emph{features}. In particular, let $f_i = [f_{i1}, \, f_{i2}, \ldots]$ denote a binary feature vector for the $i^{th}$ time series. Setting $f_{ik}=1$ implies that time series $i$ exhibits behavior~$k$ for some subset of values $t \in \{1,\dots,T_i\}$, where $T_i$ is the length of the $i^{th}$ time series. Our proposed featural model defines $N$ such infinite-dimensional feature vectors, one for each time series. By discovering the pattern of behavior-sharing via a featural model (i.e., discovering $f_{ik}=f_{jk}=1$ for some $i,j,k$), we can interpret how the time series relate to one another in addition to harnessing the shared structure to pool observations from the same behavior, thus improving our estimate of $\theta_k$.
\subsection{Beta Processes}
\label{background:BetaProcess}
Inferring the structure of behavior sharing within a Bayesian framework requires defining a prior on the feature inclusion probabilities. Since we want to maintain an unbounded set of possible behaviors (and thus require infinite-dimensional feature vectors), we appeal to a Bayesian nonparametric featural model based on the \emph{beta process-Bernoulli process}. Informally, one can think of the beta process as defining an infinite set of coin-flipping probabilities and each Bernoulli process realization is the outcome from an infinite coin-flipping sequence based on the beta-process-determined coin weights. The set of resulting \emph{heads} indicate the set of selected \emph{features}, and implicitly defines an infinite-dimensional \emph{feature vector}. The properties of the beta process induce sparsity in the feature space by encouraging sharing of features among the Bernoulli process realizations. The inherent conjugacy of the beta process to the Bernoulli process allows for an analytic predictive distribution on a feature vector (i.e., Bernoulli realization) based on the feature vectors observed so far (i.e., previous Bernoulli process draws). As outlined in Section~\ref{sec:IBP}, this predictive distribution can be described via the Indian buffet process under certain parameterizations of the beta process.
\subsubsection*{The Beta Process - Bernoulli Process Featural Model}
The beta process is a special case of a general class of stochastic processes known as~\emph{completely random measures}~\citep{Kin1967}. A completely random measure $B$ is defined such that for any disjoint sets $A_1$ and $A_2$ (of some sigma algebra $\mathcal{A}$ on a measurable space $\Theta$), the corresponding random measures $B(A_1)$ and $B(A_2)$ are independent. This idea generalizes the family of \emph{independent increments processes} on the real line. All completely random measures can be constructed from realizations of a nonhomogenous Poisson process (up to a deterministic component)~\citep{Kin1967}. Specifically, a Poisson rate measure $\eta$ is defined on a product space $\Theta \otimes \mathbb{R}$, and a draw from the specified Poisson process yields a collection of points $\{\theta_j,\omega_j\}$ that can be used to define a completely random measure:
\begin{align}
B = \sum_{k=1}^\infty \omega_k\delta_{\theta_k}.
\label{eqn:CRM}
\end{align}
This construction assumes $\eta$ has infinite mass, yielding the countably infinite collection of points from the Poisson process. From Eq.~\eqref{eqn:CRM}, we see that completely random measures are discrete. Consider a rate measure defined as the product of an arbitrary sigma-finite \emph{base measure} $B_0$, with total mass $B_0(\Theta)=\alpha$, and an improper beta distribution on the product space $\Theta \otimes [0,1]$:
\begin{equation}
\nu(d\omega, d\theta) = c\omega^{-1}(1 - \omega)^{c-1}d\omega B_0(d\theta),
\end{equation}
where $c>0$ is referred to as a \emph{concentration parameter}. The resulting completely random measure is known as the \emph{beta process} with draws denoted by $B \sim \mbox{BP}(c,B_0)$~\footnote{Letting the rate measure be defined as a product of a base measure $G_0$ and an improper gamma distribution $\eta(d\theta,d\omega) = cp^{-1}e^{-cp}dpG_0(d\theta)$, with $c>0$, gives rise to completely random measures $G\sim \mbox{GP}(c,G_0)$, where GP denotes a \emph{gamma process}. Normalizing $G$ yields draws from a Dirichlet process $\DP{\alpha}{G_0/\alpha}$, with $\alpha = G_0(\Theta)$. Note that these random \emph{probability} measures $G$ are necessarily not completely random since the random variables $G(A_1)$ and $G(A_2)$ for disjoint sets $A_1$ and $A_2$ are dependent due to the normalization constraint.}. Note that using this construction, the weights $\omega_k$ of the atoms in $B$ lie in the interval $(0,1)$. Since $\eta$ is $\sigma$-finite, Campbell's theorem~\citep{Kingman:93} guarantees that for $\alpha$ finite, $B$ has finite expected measure. For an example realization and its associated cumulative distribution, see Fig.~\ref{fig:BPBePrealizations}.
Note that for a base measure $B_0$ containing atoms, a sample $B \sim \mbox{BP}(c,B_0)$ necessarily contains each of these atoms $\theta_k$ with associated weights
\begin{align}
\omega_k \sim \mbox{Beta}(c q_k, c(1-q_k)),
\end{align}
where $q_k \in (0,1)$ denotes the mass of the $k^{th}$ atom in $B_0$.
\begin{figure}[t]
\centering
\begin{tabular}{cc}
\includegraphics[width=0.5\columnwidth]{figs/BP_BeP_realizations}
& \includegraphics[width =
0.4\columnwidth]{figs/F_sample_from_prior_gamma10}\\
(a) & (b)
\end{tabular}
\caption[A draw from a beta process, and associated
Bernoulli realizations, along with a realization from the Indian
buffet process.]{(a) \textit{Top}: A draw $B$ from a beta process is
shown in blue, with the corresponding cumulative distribution in
red. \textit{Bottom}: 50 draws $X_i$ from a Bernoulli process using
the beta process realization. Each blue dot corresponds to a
coin-flip at that atom in $B$ that came up heads. (b) An image of a
feature matrix associated with a realization from an Indian buffet
process with $\alpha=10$. Each row corresponding to a different
customer, and each column a different dish. White indicates a
chosen feature.} \label{fig:BPBePrealizations}
\end{figure}
The beta process is conjugate to a class of \emph{Bernoulli processes}~\citep{Thibaux:07}, denoted by $\mbox{BeP}(B)$, which provide our sought-for featural representation. A realization
\begin{align}
X_i\mid B \sim \mbox{BeP}(B),
\end{align}
with $B$ an atomic measure, is a collection of unit-mass atoms on $\Theta$ located at some subset of the atoms in $B$. In particular,
\begin{align}
f_{ik} \sim \mbox{Bernoulli}(\omega_k)
\end{align}
is sampled independently for each atom $\theta_k$ in $B$~\footnote{One can visualize this process as walking along the atoms of a discrete measure $B$ and, at each atom $\theta_k$, flipping a coin with probability of heads given by $\omega_k$.}, and then
\begin{align}
X_i = \sum_k f_{ik} \delta_{\theta_k}.
\end{align}
Example realizations of $X_i \sim \mbox{BeP}(B)$, with $B$ a draw from a beta process, are shown in Fig.~\ref{fig:BPBePrealizations}(a).
For continuous measures $B$, we draw $L \sim \mbox{Poisson}(B(\Theta))$ and then independently sample a set of $L$ atoms $\theta_\ell \sim B(\Theta)^{-1}B$. The Bernoulli realization is then given by:
\begin{align}
X_i = \sum_{\ell=1}^L \delta_{\theta_\ell}.
\end{align}
In our subsequent development, we interpret the atom locations $\theta_k$ as a set of global features that can be shared among multiple time series. A Bernoulli process realization $X_i$ then determines the subset of features allocated to time series~$i$:
\begin{align}
B \mid B_0, c&\sim \mbox{BP}(c,B_0)\nonumber\\
X_i \mid B &\sim \mbox{BeP}(B), \quad i=1,\dots,N. \label{eqn:HierarchicalBeta}
\end{align}
Computationally, Bernoulli process realizations $X_i$ are often summarized by an infinite vector of binary indicator variables $\fset{i} = [f_{i1}, f_{i2}, \ldots]$, where $f_{ik}=1$ if and only if time series~$i$ exhibits feature~$k$. Using the beta process measure $B$ to tie together the feature vectors encourages them to share similar features while allowing time-series-specific variability.
\section{Describing Multiple Time Series with Beta Processes} \label{sec:model}
We employ the beta process featural model of Section~\ref{background:BetaProcess} to define a prior on the collection of \emph{infinite}-dimensional feature vectors $\fset{i} = [f_{i1}, \, f_{i2}, \ldots]$ used to describe the relationship amongst our $N$ time series. Recall from Section~\ref{sec:multipleTimeSeries} that the globally-shared parameters $\theta_k$ define the possible \emph{behaviors} (e.g., VAR processes), while the feature vector $\fset{i}$ indicates the behaviors exhibited by time series $i$.
\subsubsection*{Beta Process Prior on Features}
In our scenario, the beta process hierarchy of Equation~\eqref{eqn:HierarchicalBeta} can be interpreted as follows. The random measure $B \sim \mbox{BP}(c,B_0)$ defines a set of weights on the global collection of behaviors. Then, each \emph{time series} $i$ is associated with a draw from a Bernoulli process, $X_i\mid B \sim \mbox{BeP}(B)$. The Bernoulli process realization $X_i = \sum_k f_{ik} \delta_{\theta_k}$ implicitly defines the feature vector $\fset{i}$ for time series $i$, indicating which set of globally-shared behaviors that time series has selected. Such a featural model seeks to allow for infinitely many possible behaviors, while encouraging a sparse, finite representation and flexible sharing of behaviors between time series. For example, the lower subfigure in Fig.~\ref{fig:BPBePrealizations}(a) illustrates a collection of feature vectors drawn from this process.
Conditioned on the set of $N$ feature vectors $\fset{i}, i=1,\dots,N$ drawn from the hierarchy of Equation~\eqref{eqn:HierarchicalBeta}, the model reduces to a collection of $N$ switching VAR processes, each defined on the finite state space formed by the set of selected behaviors for that time series. In the following section, we define the generative process for the Markov dynamics based on the sampled feature vectors.
\subsubsection*{Feature-Constrained Transition Distributions}
Given $\fset{i}$, the $i^{th}$ time series's Markov transitions among its set of dynamic behaviors are governed by a set of \emph{feature-constrained transition distributions} \mbox{$\BF{\pi}^{(i)} =\{\symsubsup{\pi}{k}{i}\}$}. In particular, motivated by the fact that Dirichlet-distributed probability mass functions can be generated via normalized gamma random variables, for each time series $i$ we define a doubly infinite collection of random variables:
\begin{equation}
\symsubsup{\eta}{jk}{i}\mid \gamma,\kappa \sim \mbox{Gamma}(\gamma+\kappa\delta(j,k),1), \label{eqn:transitionGamma}
\end{equation}
Here, $\delta(j,k)$ indicates the Kronecker delta function. Using this collection of \emph{transition variables}, denoted by $\BF{\eta}^{(i)}$, one can define time-series-specific, feature-constrained transition distributions:
\begin{align}
\symsubsup{\pi}{j}{i} = \frac{
\begin{bmatrix}
\symsubsup{\eta}{j1}{i} & \symsubsup{\eta}{j2}{i} & \dots\;
\end{bmatrix}
\otimes \fset{i}}{\sum_{k|f_{ik}=1} \symsubsup{\eta}{jk}{i}}, \label{eqn:normEta}
\end{align}
where $\otimes$ denotes the element-wise, or Hadamard, vector product. This construction defines $\symsubsup{\pi}{j}{i}$ over the full set of positive integers, but assigns positive mass only at indices~$k$ where $f_{ik}=1$, thus constraining time series $i$ to solely transition amongst the dynamical behaviors indicated by its feature vector $\fset{i}$.
The preceding generative process can be equivalently represented via a sample $\symsubsup{\tilde{\pi}}{j}{i}$ from a finite Dirichlet distribution of dimension $K_i = \sum_k f_{ik}$, containing the non-zero entries of $\symsubsup{\pi}{j}{i}$:
\begin{equation}
\symsubsup{\tilde{\pi}}{j}{i} \mid \fset{i}, \gamma,\kappa \sim \mbox{Dir}([\gamma,\dots,\gamma,\gamma + \kappa,\gamma,\dots\gamma]). \label{eqn:DirPrior}
\end{equation}
The $\kappa$ hyperparameter places extra expected mass on the component of $\symsubsup{\tilde{\pi}}{j}{i}$ corresponding to a self-transition $\symsubsup{\pi}{jj}{i}$, analogously to the sticky hyperparameter of the sticky HDP-HMM~\citep{Fox:AOAS11}. We also use the representation
\begin{align}
\symsubsup{\pi}{j}{i} \mid \fset{i},\gamma,\kappa \sim \mbox{Dir}([\gamma,\dots,\gamma,\gamma + \kappa,\gamma,\dots]\otimes \fset{i}),\label{eqn:DirPrior2}
\end{align}
implying $\symsubsup{\pi}{j}{i} =
\begin{bmatrix}
\symsubsup{\pi}{j1}{i} & \symsubsup{\pi}{j2}{i} & \dots
\end{bmatrix}$, with only a finite number of non-zero entries $\symsubsup{\pi}{jk}{i}$. This representation is really an abuse of notation since the Dirichlet distribution is not defined for infinitely many parameters. In reality, we are simply examining a $K_i$-dimensional Dirichlet distribution as in Eq.~\eqref{eqn:DirPrior}. However, the notation of Eq.~\eqref{eqn:DirPrior2} is useful in reminding the reader that the indices of $\symsubsup{\tilde{\pi}}{j}{i}$ defined by Eq.~\eqref{eqn:DirPrior} are not over 1 to $K_i$, but rather over the $K_i$ values of $k$ such that $f_{ik}=1$. Additionally, this notation is useful for concise representations of the posterior distribution.
\subsubsection*{VAR Likelihoods}
Although the methodology described thus far applies equally well to HMMs and other Markov switching processes, henceforth we focus our attention on the AR-HMM and develop the full model specification and inference procedures needed to treat our motivating example of visual motion capture. Specifically, let $\symsubsupB{y}{t}{i}$ represent the observed value of the $i^{th}$ time series at time $t$, and let $\symsubsup{z}{t}{i}$ denote the latent dynamical mode. Assuming an order-$r$ AR-HMM, we have
\begin{equation}
\begin{aligned}
\symsubsup{z}{t}{i} &\sim \symsubsup{\pi}{\symsubsup{z}{t-1}{i}}{i}\\
\symsubsupB{y}{t}{i} &= \sum_{j=1}^r A_{j,\symsubsup{z}{t}{i}}\symsubsupB{y}{t-j}{i} + \symsubsupB{e}{t}{i}(\symsubsup{z}{t}{i}) \triangleq \BF{A}_{\symsubsup{z}{t}{i}}\symsubsupB{\tilde{y}}{t}{i} + \symsubsupB{e}{t}{i}(\symsubsup{z}{t}{i}),
\end{aligned}
\label{eqn:multSVAR}
\end{equation}
where $\symsubsupB{e}{t}{i}(k) \sim \mathcal{N}(0,\Sigma_{k})$, $\BF{A}_{k} =
\begin{bmatrix}
A_{1,k} & \dots & A_{r,k}
\end{bmatrix}
$, and $\symsubsupB{\tilde{y}}{t}{i} =
\begin{bmatrix}
\smash{\symsubsupT{y}{t-1}{i}} & \dots & \smash{\symsubsupT{y}{t-r}{i}}
\end{bmatrix}^T$. Recall that each of the behaviors $\theta_k = \{\BF{A}_{k},\Sigma_k\}$ defines a different VAR($r$) dynamical mode and the feature-constrained transition distributions $\pi^{(i)}$ restrict time series $i$ to select among dynamic behaviors (indexed at time $t$ by $\symsubsup{z}{t}{i}$) that were picked out by its feature vector $\fset{i}$. Our beta-process-based featural model couples the dynamic behaviors exhibited by different time series.
\subsubsection*{Prior on VAR Parameters}
To complete the Bayesian model specification, a conjugate matrix-normal inverse-Wishart (MNIW) prior (cf.,~\cite{West}) is placed on the shared collection of dynamic parameters $\theta_k = \{\BF{A}_k,\Sigma_k\}$. Specifically, this prior is comprised of an inverse Wishart prior on $\Sigma_k$ and (conditionally) a matrix normal prior on $\BF{A}_k$:
\begin{equation}
\begin{aligned}
\Sigma_k \mid n_0,S_0 &\sim \mbox{IW}(n_0,S_0)\\
\BF{A}_k \mid \Sigma_k,M,K &\sim \MN{\BF{A}_k}{M}{\Sigma_k}{K},
\end{aligned}
\end{equation}
with $n_0$ the degrees of freedom, $S_0$ the scale matrix, $M$ the mean dynamic matrix, and $K$ a covariance matrix that together with $\Sigma_k$ defines the covariance of $A_k$. This prior defines the base measure $B_0$ up to the total mass parameter $\alpha$, which has to be separately assigned. As motivated in Section~\ref{sec:IBPhyperparameters}, this latter parameter is given a gamma prior.
Since the library of possible dynamic parameters is shared by all time series, posterior inference of each parameter set $\theta_k$ relies on pooling data amongst the time series that have $f_{ik}=1$. It is through this pooling of data that one may achieve more robust parameter estimates than from considering each time series individually.
\subsubsection*{The BP-AR-HMM}
We term the resulting model the \emph{BP-autoregressive-HMM} (BP-AR-HMM), with a graphical model representation presented in Fig.~\ref{fig:BPARHMM}. Considering the \emph{feature space} (i.e., set of autoregressive parameters) and the \emph{temporal dynamics} (i.e., set of transition distributions) as separate dimensions, one can think of the BP-AR-HMM as a spatio-temporal process comprised of a (continuous) beta process in space and discrete-time Markovian dynamics in time. The overall model specification is summarized as:
\begin{equation}
\begin{aligned}
B \mid B_0 &\sim \mbox{BP}(1,B_0)\\
X_i \mid B &\sim \mbox{BeP}(B), \quad i = 1,\dots, N\\
\symsubsup{\pi}{j}{i} \mid \fset{i},\gamma,\kappa &\sim \mbox{Dir}([\gamma,\dots,\gamma,\gamma + \kappa,\gamma,\dots]\otimes \fset{i}), \quad i=1,\dots,N, \,\, j=1,2,\dots\\
\symsubsup{z}{t}{i} &\sim \symsubsup{\pi}{\symsubsup{z}{t-1}{i}}{i}, \quad i=1,\dots,N, \,\, t=1,\dots,T_i\\
\symsubsupB{y}{t}{i} &= \BF{A}_{\symsubsup{z}{t}{i}}\symsubsupB{\tilde{y}}{t}{i} + \symsubsupB{e}{t}{i}(\symsubsup{z}{t}{i}), \quad i=1,\dots,N, \,\, t=1,\dots,T_i.
\end{aligned}
\label{eqn:BPARHMM}
\end{equation}
\begin{figure}
[t!] \centering \hspace{0.2in}
\includegraphics[height=2.5in]{figs/IBPARHMM3}\vspace{-0.2in} \caption[Graphical model of the BP-AR-HMM.] {Graphical model of the BP-AR-HMM. The beta process distributed measure $\mbox{$B \mid B_0 \sim \mbox{BP}(1,B_0)$}$ is represented by its masses $\omega_k$ and locations $\theta_k$, as in Eq.~\eqref{eqn:CRM}. The features are then conditionally independent draws $\mbox{$f_{ik} \mid \omega_k \sim \mbox{Bernoulli}(\omega_k)$}$, and are used to define feature-constrained transition distributions $\mbox{$\symsubsup{\pi}{j}{i} \mid \fset{i}, \gamma,\kappa \sim \mbox{Dir}([\gamma,\dots,\gamma,\gamma+\kappa,\gamma,\dots]\otimes \fset{i})$}$. The switching VAR dynamics are as in Eq.~\eqref{eqn:multSVAR}.} \label{fig:BPARHMM}
\end{figure}
\section{MCMC Posterior Computations} \label{sec:MCMC}
In this section, we develop an MCMC method which alternates between sampling binary feature assignments given observations and dynamic parameters, and sampling dynamic parameters given observations and features. The sampler interleaves Metropolis-Hastings and Gibbs sampling updates, which are sometimes simplified by appropriate auxiliary variables. We leverage the fact that fixed feature assignments instantiate a set of \emph{finite} AR-HMMs, for which dynamic programming can be used to efficiently compute marginal likelihoods. Computationally, sampling the potentially infinite set of time-series-specific features in our beta process featural model relies on a predictive distribution on features that can be described via the \emph{Indian buffet process} (IBP)~\citep{GriffithsGhahramani:05}. The details of the IBP are outlined below. As a key component of our feature-sampling, we introduce a new approach employing incremental ``birth'' and ``death'' proposals, improving on previous exact samplers for IBP models in the non-conjugate case~\citep{Meeds:07}.
\subsection{Background: The Indian Buffet Process}
\label{sec:IBP}
As shown by~\citet{Thibaux:07}, marginalizing over the latent beta process $B$ in the hierarchical model of Equation~\eqref{eqn:HierarchicalBeta}, and taking $c=1$, induces a predictive distribution on feature indicators known as the Indian buffet process (IBP)~\citep{GriffithsGhahramani:05}. The IBP is based on a culinary metaphor in which customers arrive at an infinitely long buffet line of dishes, or features (\emph{behaviors} in our case). The first arriving customer, or \emph{time series} in our case, chooses $\mbox{Poisson}(\alpha)$ dishes. Each subsequent customer~$i$ selects a previously tasted dish~$k$ with probability $m_k/i$ proportional to the number of previous customers $m_k$ to sample it, and also samples $\mbox{Poisson}(\alpha/i)$ new dishes. The feature matrix associated with a realization from an Indian buffet process is shown in Fig.~\ref{fig:BPBePrealizations}(b).
To derive the IBP from the beta process formulation described above,
we note that the probability $X_{i}$ contains feature $\theta_k$
after having observed $X_1,\dots,X_{i-1}$ is equal to the expected
mass of that atom:
\begin{align}
p(f_{ik}=1 \mid X_1,\dots,X_{i-1}) = \mathbb{E}_{B\mid
X_1,\dots,X_{i-1}}[ p(f_{ik}=1\mid B)] = \mathbb{E}_{B\mid
X_1,\dots,X_{i-1}}[\omega_k],
\end{align}
where our notation $\mathbb{E}_{B}[\cdot]$ means to take the
expectation with respect to the distribution of $B$. Because beta process priors are conjugate to the Bernoulli process~\citep{Kim:99}, the posterior distribution given $N$ samples $X_i \sim \mbox{BeP}(B)$ is a beta process with updated parameters:
\begin{align}
B \mid X_1,\dots,X_N, B_0, c &\sim
\mbox{BP}\Bigg(c+N,\frac{c}{c+N}B_0 + \frac{1}{c+N}\sum_{i=1}^N
X_i\Bigg)\\
&= \mbox{BP}\Bigg(c+N,\frac{c}{c+N}B_0 + \sum_{k=1}^{\ensuremath{K_{\!+}}}
\frac{m_k}{c+N}\delta_{\theta_k}\Bigg). \label{eqn:betapost}
\end{align}
Here, $m_k$ denotes the number of time series $X_i$ that select the $k^{th}$ feature $\theta_k$ (i.e., $f_{ik}=1$). For simplicity, we have reordered the feature indices to list first the $\ensuremath{K_{\!+}}$ features used by at least one time series.
Using the
posterior distribution defined in Eq.~\eqref{eqn:betapost}, we
consider the discrete and continuous portions of the base measure
separately. The discrete component is a collection of atoms at
locations $\theta_1,\dots,\theta_{K_+}$, each with weight
\begin{align}
q_k = \frac{m_k}{c+i-1},
\end{align}
where $K_+$ is the number of unique atoms present in
$X_1,\dots,X_{i-1}$. For each of the currently instantiated features
$k \in \{1,\dots,K_+\}$, we have
\begin{align}
\omega_k \sim \mbox{Beta}((c+i-1)q_k,(c+i-1)(1-q_k))
\end{align}
such that the expected weight is simply $q_k$, implying that the
$i^{th}$ time series chooses one of the currently instantiated features
with probability proportional to the number of time series that already
chose that feature, $m_k$. We now consider the continuous portion of
the base measure,
\begin{align}
\frac{c}{c+i-1}B_0.
\end{align}
The Poisson process defined by this rate function generates
\begin{align}
\mbox{Poisson}\left(\frac{c}{c+i-1}B_0(\Theta)\right) =
\mbox{Poisson}\left(\frac{c}{c+i-1}\alpha\right)
\end{align}
new atoms in $X_i$ that do not appear in $X_1,\dots,X_{i-1}$.
Following this argument, the first time series simply chooses
$\mbox{Poisson}(\alpha)$ features. If we specialize this process to
$c=1$, we arrive at the IBP.
\subsection{Sampling binary feature assignments} \label{sec:featureSampling}
Let $\BF{F}^{-ik}$ denote the set of all binary feature indicators excluding $f_{ik}$, and $K_+^{-i}$
be the number of behaviors used by all of the other time series~\footnote{Some of the $K_+^{-i}$ features may also be used by time series $i$, but only those not unique to that time series.}. For notational simplicity, we assume that these behaviors are indexed by $\{1,\dots,K_+^{-i}\}$. The IBP prior differentiates between features, or behaviors, that other time series have already selected and those unique to the current time series. Thus, we examine each of these cases separately.
\subsubsection*{Shared features} Given the $i^{th}$ time series $\symsubsupB{y}{1:T_i}{i}$, transition variables $\BF{\eta}^{(i)} = \symsubsup{\eta}{1:K_+^{-i},1:K_+^{-i}}{i}$, and shared dynamic parameters $\theta_{1:K_+^{-i}}$, the feature indicators $f_{ik}$ for currently used features $k \in \{1,\dots,K_+^{-i}\}$ have the following posterior distribution:
\begin{equation}
p(f_{ik}\mid \BF{F}^{-ik}\!,\symsubsupB{y}{1:T_i}{i},\BF{\eta}^{(i)}\!, \theta_{1:K_+^{-i}},\alpha) \propto p(f_{ik}\mid \BF{F}^{-ik}\!, \alpha) p(\symsubsupB{y}{1:T_i}{i}\mid \fset{i}, \BF{\eta}^{(i)}\!,\theta_{1:K_+^{-i}}). \label{eqn:Fsampling}
\end{equation}
Here, the IBP prior described in Section~\ref{background:BetaProcess} implies that $p(f_{ik}=1\mid \BF{F}^{-ik}\!, \alpha) = m_k^{-i}/N$, where $m_k^{-i}$ denotes the number of time series \emph{other} than time series $i$ that exhibit behavior $k$. In evaluating this expression, we have exploited the exchangeability of the IBP~\citep{GriffithsGhahramani:05}, which follows directly from the beta process construction~\citep{Thibaux:07}.
For binary random variables, Metropolis-Hastings proposals can mix faster~\citep{Frigessi:93} and have greater efficiency~\citep{Liu:96} than standard Gibbs samplers. To update $f_{ik}$ given $\BF{F}^{-ik}\!$, we thus use the posterior of Eq.~\eqref{eqn:Fsampling} to evaluate a Metropolis-Hastings proposal which flips $f_{ik}$ to the complement $\bar{f}$ of its current value $f$:
\begin{align}
f_{ik} &\sim \rho(\bar{f} \mid f)\delta(f_{ik},\bar{f}) + (1-\rho(\bar{f} \mid f))\delta(f_{ik},f) \nonumber \\
\rho(\bar{f} \mid f) &= \min \Bigg\{\frac{p(f_{ik}=\bar{f}\mid \BF{F}^{-ik}\!,\symsubsupB{y}{1:T_i}{i},\BF{\eta}^{(i)}\!,\theta_{1:K_+^{-i}},\alpha)}{p(f_{ik}=f\mid \BF{F}^{-ik}\!,\symsubsupB{y}{1:T_i}{i},\BF{\eta}^{(i)}\!,\theta_{1:K_+^{-i}},\alpha)},1\Bigg\}. \label{eqn:sharedFeaturesMH}
\end{align}
To compute likelihoods, we combine $\fset{i}$ and $\BF{\eta}^{(i)}$ to construct feature-constrained transition distributions $\symsubsup{\pi}{j}{i}$ as in Eq.~\eqref{eqn:normEta}, and marginalize over the exponentially large set of possible latent mode sequences by applying a variant of the sum-product message passing algorithm for AR-HMMs. (See Appendix~\ref{app:sumprod}.)
\subsubsection*{Unique features} An alternative approach is needed to sample the $\mbox{Poisson}(\alpha/N)$ ``unique'' features associated only with time series~$i$. Let $K_+ = K_+^{-i}+
\newfeatures{i}$, where $
\newfeatures{i}$ is the number of unique features chosen, and define $\fset{-i} = f_{i,1:K_+^{-i}}$ and $\fset{+i} = f_{i,K_+^{-i}+1:K_+}$. The posterior distribution over $
\newfeatures{i}$ is then given by
\vspace{-0.1in}
\begin{multline}
p(
\newfeatures{i} \mid \fset{i},\symsubsupB{y}{1:T_i}{i},\BF{\eta}^{(i)}\!,\theta_{1:K_+^{-i}},\alpha) \propto \frac{(\frac{\alpha}{N})^{
\newfeatures{i}}e^{-\frac{\alpha}{N}}}{
\newfeatures{i}!}\\
\iint p(\symsubsupB{y}{1:T_i}{i}\mid \fset{-i},\fset{+i}=\BF{1}, \BF{\eta}^{(i)}\!,\BF{\eta}_+,\theta_{1:K_+^{-i}},\BF{\theta}_+) \;dB_0(\BF{\theta}_+)dH(\BF{\eta}_+),
\end{multline}
where $H$ is the gamma prior on transition variables $\symsubsup{\eta}{jk}{i}$, and we recall that $B_0$ is the base measure of the beta process. The set $\BF{\theta}_+ = \theta_{K_+^{-i} + 1: K_+}$ consists of the parameters of unique features, and $\BF{\eta}_+$ the transition parameters $\symsubsup{\eta}{jk}{i}$ to or from unique features $j,k \in \{K_+^{-i} + 1:K_+\}$. Exact evaluation of this integral is intractable due to dependencies induced by the AR-HMMs.
One early approach to approximate Gibbs sampling in non-conjugate IBP models relies on a finite truncation of the limiting Bernoulli process~\citep{Gorur:06}. That is, drawing $n_i \sim \mbox{Poisson}(\alpha/N)$ distribution is equivalent to setting $n_i$ equal to the number of successes in infinitely many Bernoulli trials, each with probability of success
\begin{align}
\lim_{K \rightarrow \infty} \frac{\alpha/K}{\alpha/K + N}.
\end{align}
\citet{Gorur:06} truncate this process and instead consider $K^*$ Bernoulli trials with probability $(\alpha/K^*)/(\alpha/K^* + N)$. \citet{Meeds:07} instead consider independent Metropolis proposals which replace the existing unique features by $n_i \sim \mbox{Poisson}(\alpha/N)$ new features, with corresponding parameters $\BF{\theta}_+$ drawn from the prior. For high-dimensional models such as those considered in this paper, however, such moves have extremely low acceptance rates.
We instead develop a birth and death reversible jump MCMC sampler~\citep{Green:95}, which proposes to either add a single new feature, or eliminate one of the existing features in $\fset{+i}$. Our proposal distribution factors as follows:
\begin{equation}
q(\fset{+i}',\BF{\theta}_{+}',\BF{\eta}_{+}' \mid \fset{+i},\BF{\theta}_{+},\BF{\eta}_{+}) = q_f(\fset{+i}' \mid \fset{+i}) q_{\theta}(\BF{\theta}_{+}' \mid \fset{+i}',\fset{+i},\BF{\theta}_{+}) q_{\eta}(\BF{\eta}_{+}' \mid \fset{+i}', \fset{+i}, \BF{\eta}_{+})\label{eqn:jointproposal}
\end{equation}
Let $n_i = \sum_k\! f_{+ik}$. The feature proposal $q_f(\cdot\mid\cdot)$ encodes the probabilities of birth and death moves, which we set as follows: A new feature is created with probability $0.5$, and each of the $n_i$ existing features is deleted with probability $0.5/n_i$. This set of possible proposals leads to considering transitions from $n_i$ to $n_i'$ unique features, with $n_i'=n_i+1$ in the case of a birth proposal, or $n_i'=n_i-1$ in the case of a proposed feature death. Note that if the proposal from the distribution defined in Eq.~\eqref{eqn:jointproposal} is rejected, we maintain $n_i' = n_i$ unique features. For parameters, we define our proposal using the generative model:
\begin{align}
q_{\theta}(\BF{\theta}_{+}' \mid \fset{+i}',\fset{+i},\BF{\theta}_{+}) = \left\{
\begin{array}{ll}
b_0(\theta_{+,n_i+1}') \prod_{k=1}^{n_i} \delta_{\theta_{+,k}}(\theta_{+,k}'), & \hbox{birth of feature } n_i + 1; \\
\prod_{k\neq \ell}\delta_{\theta_{+,k}}(\theta_{+,k}'), & \hbox{death of feature } \ell.
\end{array}
\right.\label{eqn:parameter_proposals}
\end{align}
That is, for a birth proposal, a new parameter $\theta_{+,n_i+1}'$ is drawn from the prior and all other parameters remain the same. For a death proposal of feature $j$, we simply eliminate that parameter from the model. Here, $b_0$ is the density associated with $\alpha^{-1}B_0$. The distribution $q_{\eta}(\cdot\mid\cdot)$ is defined similarly, but using the gamma prior on transition variables of Eq.~\eqref{eqn:transitionGamma}.
The Metropolis-Hastings acceptance probability is then given by
\begin{equation}
\rho(\fset{+i}',\BF{\theta}_{+}',\BF{\eta}_{+}' \mid \fset{+i},\BF{\theta}_{+},\BF{\eta}_{+}) = \min\{r(\fset{+i}',\BF{\theta}_{+}',\BF{\eta}_{+}' \mid \fset{+i},\BF{\theta}_{+},\BF{\eta}_{+}),1\}.
\end{equation}
As derived in Appendix~\ref{app:birthdeath}, we compactly represent the acceptance ratio $r(\cdot \mid \cdot)$ for either a birth or death move as
\begin{equation}
\frac{p(\symsubsupB{y}{1:T_i}{i}\mid [\fset{-i} \, \fset{+i}'], \theta_{1:K_+},\BF{\theta}_{+}',\BF{\eta}^{(i)}, \BF{\eta}_{+}') \; \mbox{Poisson}(n_i' \mid \alpha/N) \; q_f(\fset{+i} \mid \fset{+i}')}{ p(\symsubsupB{y}{1:T_i}{i}\mid [\fset{-i} \, \fset{+i}], \theta_{1:K_+},\BF{\eta}^{(i)}) \; \mbox{Poisson}(n_i \mid \alpha/N) \; q_f(\fset{+i}' \mid \fset{+i})},\label{eqn:uniqueFeaturesMH}
\end{equation}
where we recall that $n_i' = \sum_k\! f_{+ik}'$. Because our birth and death proposals do not modify the values of existing parameters, the Jacobian term normally arising in reversible jump MCMC algorithms simply equals one.
\subsection{Sampling dynamic parameters and transition variables}
Posterior updates to transition variables $\BF{\eta}^{(i)}$ and shared dynamic parameters $\theta_k$ are greatly simplified if we instantiate the mode sequences $\symsubsup{z}{1:T_i}{i}$ for each time series $i$. We treat these mode sequences as auxiliary variables that are discarded for subsequent updates of feature assignments $\fset{i}$.
\subsubsection*{Mode sequences $\symsubsup{z}{1:T_i}{i}$} Given feature-constrained transition distributions $\BF{\pi}^{(i)}$ and dynamic parameters $\{\theta_k\}$, along with the observation sequence $\symsubsupB{y}{1:T_i}{i}$, we block sample the mode sequence $\symsubsup{z}{1:T_i}{i}$ by computing backward messages $m_{t+1,t}(\symsubsup{z}{t}{i}) \propto p(\symsubsupB{y}{t+1:T_i}{i} \mid \symsubsup{z}{t}{i},\symsubsupB{\tilde{y}}{t}{i},\BF{\pi}^{(i)},\{\theta_k\})$, and then recursively sampling each $\symsubsup{z}{t}{i}$:
\begin{equation}
\symsubsup{z}{t}{i} \mid \symsubsup{z}{t-1}{i}, \symsubsupB{y}{1:T_i}{i},\BF{\pi}^{(i)}\!, \{\theta_k\} \sim \symsubsup{\pi}{\symsubsup{z}{t-1}{i}}{i}\!(\symsubsup{z}{t}{i}) \mathcal{N}\big(\symsubsupB{y}{t}{i}; \BF{A}_{\symsubsup{z}{t}{i}}\symsubsupB{\tilde{y}}{t}{i}, \Sigma_{\symsubsup{z}{t}{i}}\big) m_{t+1,t}(\symsubsup{z}{t}{i}).
\end{equation}
This backward message-passing, forward-sampling scheme is detailed in Appendix~\ref{app:sumprod}.
\subsubsection*{Transition distributions $\symsubsup{\pi}{j}{i}$}
We use the fact that Dirichlet priors are conjugate to multinomial observations $\symsubsup{z}{1:T}{i}$ to derive the posterior of $\symsubsup{\pi}{j}{i}$ as
\begin{align}
\symsubsup{\pi}{j}{i} \mid \fset{i},\symsubsup{z}{1:T}{i}, \gamma,\kappa \sim \mbox{Dir}([\gamma+\symsubsup{n}{j1}{i},\dots,\gamma + \symsubsup{n}{jj-1}{i},\gamma + \kappa + \symsubsup{n}{jj}{i},\gamma + \symsubsup{n}{jj+1}{i},\dots]\otimes \fset{i}).\label{eqn:piPosterior}
\end{align}
Here, $\symsubsup{n}{jk}{i}$ are the number of transitions from mode $j$ to $k$ in $\symsubsup{z}{1:T}{i}$. Since the mode sequence $\symsubsup{z}{1:T}{i}$ was generated from feature-constrained transition distributions, $\symsubsup{n}{jk}{i}$ will be zero for any $k$ such that $f_{ik}=0$. Using the definition of $\symsubsup{\pi}{j}{i}$ in Eq.~\eqref{eqn:normEta}, one can equivalently define a sample from the posterior of Eq.~\eqref{eqn:piPosterior} by solely updating $\symsubsup{\eta}{jk}{i}$ for instantiated features:
\begin{align}
\symsubsup{\eta}{jk}{i}\mid \symsubsup{z}{1:T}{i},\gamma,\kappa \sim \mbox{Gamma}(\gamma+\kappa\delta(j,k)+\symsubsup{n}{jk}{i},1), \quad k \in \{\ell \mid f_{i\ell}=1\}.
\end{align}
\subsubsection*{Dynamic parameters $\{\BF{A}_k,\Sigma_k\}$} We now turn to posterior updates for dynamic parameters. Recall the conjugate matrix normal inverse-Wishart (MNIW) prior on $\{\BF{A}_k,\Sigma_k\}$, comprised of an inverse-Wishart prior $\mbox{IW}(n_0,S_0)$ on $\Sigma_k$ and a matrix-normal prior $\MN{\BF{A}_k}{M}{\Sigma_k}{K}$ on $\BF{A}_k$ given $\Sigma_k$. We consider the following sufficient statistics based on the sets $\BF{Y}_{\!k} = \{\symsubsupB{y}{t}{i} \mid \symsubsup{z}{t}{i} = k, \, i=1,\ldots,N\}$ and \mbox{$\BF{\tilde{Y}}_{\!k} = \{\symsubsupB{\tilde{y}}{t}{i} \mid \symsubsup{z}{t}{i} = k, \, i=1,\ldots,N\}$} of observations and lagged observations, respectively, associated with behavior $k$:
\begin{equation}
\begin{aligned}
\symsubsup{S}{\tilde{y}\tilde{y}}{k} = \sum_{(t,i)\mid \symsubsup{z}{t}{i} = k} \symsubsupB{\tilde{y}}{t}{i}\symsubsupT{\tilde{y}}{t}{i} + \BF{K} &\hspace{0.25in} \symsubsup{S}{y\tilde{y}}{k} = \sum_{(t,i)\mid \symsubsup{z}{t}{i} = k} \symsubsupB{y}{t}{i}\symsubsupT{\tilde{y}}{t}{i} + \BF{M}\BF{K}\\
\symsubsup{S}{yy}{k} = \sum_{(t,i)\mid \symsubsup{z}{t}{i} = k} \symsubsupB{y}{t}{i}\symsubsupT{y}{t}{i} + \BF{M}\BF{K}\BF{M}^T &\hspace{0.25in} \symsubsup{S}{y|\tilde{y}}{k} = \symsubsup{S}{yy}{k} - \symsubsup{S}{y\tilde{y}}{k}S_{\tilde{y}\tilde{y}}^{-(k)}S_{\tilde{y}\tilde{y}}^{(k)^T}.
\end{aligned}
\label{eqn:Sk}
\end{equation}
It is through this pooling of data from multiple time series that we improve our inferences on shared behaviors, especially in the presence of limited data. Using standard MNIW conjugacy results, the posterior can be shown to equal
\begin{equation}
\begin{aligned}
\BF{A}_k \mid \Sigma_k,\BF{Y}_{\!k} &\sim \MN{\BF{A}_k}{\symsubsup{S}{y\tilde{y}}{k}S_{\tilde{y}\tilde{y}}^{-(k)}}{\Sigma_k}{\symsubsup{S}{\tilde{y}\tilde{y}}{k}}\\
\Sigma_k \mid \BF{Y}_{\!k} &\sim \mbox{IW}\left(|\BF{Y}_{\!k}| + n_0, \symsubsup{S}{y|\tilde{y}}{k} + S_0\right).
\end{aligned}
\end{equation}
\subsection{Sampling the BP and Dirichlet transition hyperparameters} \label{sec:IBPhyperparameters}
We additionally place priors on the Dirichlet hyperparameters $\gamma$ and $\kappa$, as well as the BP parameter $\alpha$.
\subsubsection*{BP hyperparameter $\alpha$} Let $\BF{F}=\{\BF{f}_i\}$. As derived by~\citet{GriffithsGhahramani:05}, $p(\BF{F} \mid \alpha)$ can be expressed as
\begin{align}
p(\BF{F} \mid \alpha) \propto \alpha^{K_+}\exp\bigg(-\alpha\sum_{n=1}^N \frac{1}{n}\bigg),
\end{align}
where, as before, $K_+$ is the number of unique features activated in $\BF{F}$. As in~\citet{Gorur:06}, we place a conjugate $\mbox{Gamma}(a_\alpha,b_\alpha)$ prior on $\alpha$, which leads to the following posterior distribution:
\begin{align}
p(\alpha \mid \BF{F},a_\alpha,b_\alpha) &\propto \alpha^{K_+}\exp\left(-\alpha\sum_{n=1}^N \frac{1}{n}\right) \cdot \frac{\alpha^{a_\alpha-1}\exp(-b_\alpha \alpha)}{\Gamma(\alpha)}\nonumber\\
&=\mbox{Gamma}\bigg(a_\alpha+K_+,b_\alpha+\sum_{n=1}^N \frac{1}{n}\bigg)
\end{align}
\subsubsection*{Transition hyperparameters $\gamma$ and $\kappa$} Transition hyperparameters are assigned priors $\gamma \sim \mbox{Gamma}(a_\gamma,b_\gamma)$ and $\kappa \sim \mbox{Gamma}(a_\kappa,b_\kappa)$. Because the generative process of Eq.~\eqref{eqn:transitionGamma} is non-conjugate, we rely on Metropolis-Hastings steps which iteratively sample $\gamma$ given $\kappa$, and $\kappa$ given $\gamma$. Each sub-step uses a gamma proposal distribution $q_\gamma(\cdot\mid\cdot)$ or $q_\kappa(\cdot\mid\cdot)$, respectively, with fixed variance $\sigma_\gamma^2$ or $\sigma_\kappa^2$, and mean equal to the current hyperparameter value.
As derived in Appendix~\ref{app:transparams}, the acceptance ratio for for the proposal of $\gamma$ given $\kappa$ is
\begin{align}
r(\gamma' \mid \gamma) = \frac{f(\gamma')\Gamma(\vartheta)\gamma^{\vartheta'-\vartheta-a_\gamma}}{f(\gamma)\Gamma(\vartheta')\gamma'^{\vartheta-\vartheta'-a_\gamma}} \exp\{-(\gamma' - \gamma)b_\gamma\} \sigma_\gamma^{2(\vartheta-\vartheta')},
\end{align}
where $\vartheta=\gamma^2/\sigma_{\gamma}^2$, $\vartheta'=\gamma'^2/\sigma_{\gamma}^2$, and $f(\gamma)$ is the likelihood term. Specifically, letting $\BF{\pi} = \{\pi_j^{(i)}\}$ and recalling the definition of $\symsubsup{\tilde{\pi}}{j}{i}$ from Eq.~\eqref{eqn:DirPrior} and that $K_i = \sum_k f_{ik}$, the likelihood term may be written as
\begin{align}
f(\gamma) \triangleq p(\BF{\pi}\mid \gamma,\kappa,\BF{F}) = \prod_i \prod_{k=1}^{K_i} \left\{\frac{\Gamma(\gamma K_i + \kappa)}{\left(\prod_{j=1}^{K_i-1} \Gamma(\gamma)\right)\Gamma(\gamma+\kappa)} \prod_{j=1}^{K_i} \tilde{\pi}_{kj}^{(i)^{\gamma+\kappa\delta(k,j)-1}}\right\}.
\end{align}
The Metropolis-Hastings sub-step for sampling $\kappa$ given $\gamma$ follows similarly. In this case, however, the likelihood terms simplifies to
\begin{align}
f(\kappa) \triangleq \prod_i \frac{\Gamma(\gamma K_i + \kappa)^{K_i}}{\Gamma(\gamma+\kappa)^{K_i}} \prod_{j=1}^{K_i} \tilde{\pi}_{jj}^{(i)^{\gamma+\kappa-1}} \propto p(\BF{\pi}\mid \gamma,\kappa,\BF{F}).
\end{align}
The resulting MCMC sampler for the BP-AR-HMM is summarized in Algorithm~\ref{alg:IBPARHMMsampler} of Appendix~\ref{app:alg}.
\section{Related Work}
\label{sec:related}
A challenging problem in deploying Markov switching processes such as the AR-HMM is that of defining the number of dynamic regimes. Previously, Bayesian nonparametric approaches building on the hierarchical Dirichlet process (HDP)~\citep{Teh:06} have been proposed to allow uncertainty in the number of regimes by defining Markov switching processes on infinite state spaces~\citep{Beal:02,Teh:06,Fox:AOAS11,Fox:IEEE11}. See~\citet{Fox:IEEESPM} for a recent review. However, these formulations focus on a single time series whereas in this paper our motivation is analyzing a collection of time series. A na\"{i}ve approach to employing such models in the multiple time series setting is to simply couple each of the time series under a shared HDP prior. However, such an approach assumes that the state spaces of the multiple Markov switching processes are \emph{exactly} shared, as are the transitions among these states (i.e., both the transition and emissions parameters are global.) As demonstrated in Section~\ref{sec:synth} and Section~\ref{sec:MoCap}, such strict sharing can limit the ability to discover unique dynamic behaviors and reduce the predictive performance of the inferred model.
In recent independent work, \citet{saria2010discovering} developed an alternative approach to modeling multiple time series via the HDP-HMM. Their \emph{time series topic model} (TSTM) describes coarse-scale temporal behavior using a finite set of ``topics'', which are themselves distributions on a common set of autoregressive dynamical models. Each time series is assumed to exhibit all topics to some extent, but with unique frequencies and temporal patterns.
Alternatively, the mixed HMM~\citep{altman2007mixed} uses generalized linear models to allow the state transition and emission distributions of a finite HMM to depend on arbitrary external covariates. In experiments, this is used to model the differing temporal dynamics of a small set of known time series classes.
More broadly, the specific problem we address here has received little previous attention, perhaps due to the difficulty of treating such combinatorial relationships with parametric models. There are a wide variety of models which capture correlations among multiple aligned, interacting univariate time series, for example using Gaussian state space models~\citep{aoki1991state}.
Other approaches cluster time series using a parametric mixture model~\citep{alon2003discovering}, or a Dirichlet process mixture~\citep{qi07}, and model the dynamics within each cluster via independent finite HMMs.
Dynamic Bayesian networks~\citep{murphy2002dynamic}, such as the factorial HMM~\citep{ghahramani97}, define a structured representation for the latent states underlying a single time series.
Such models are widely used in applied time series analysis~\citep{lehrach2009segmenting,duh2005jointly}.
The infinite factorial HMM~\citep{VanGael:08_2} uses the IBP to model a single time series via an infinite set of latent features, each evolving according to independent Markovian dynamics. Our work instead focuses on modeling multiple time series and on capturing dynamical modes that are shared among the series.
Other approaches do not explicitly model latent temporal dynamics, and instead aim to align time series with consistent global structure~\citep{aach2001aligning}.
Motivated by the problem of detecting temporal anomalies, \citet{listgarten2007bayesian} describe a hierarchical Bayesian approach to modeling shared structure among a known set of time series classes. Independent HMMs are used to encode non-linear alignments of observed signal traces to latent reference time series, but their states do not represent dynamic behaviors and are not shared among time series.
\section{Synthetic Experiments}
\label{sec:synth}
\subsection{Discovering Common Dynamics}
To test the ability of the BP-AR-HMM to discover shared dynamics, we generated five time series that switched between AR(1) models:
\begin{align}
\symsubsup{y}{t}{i} = a_{\symsubsup{z}{t}{i}}\symsubsup{y}{t-1}{i}+\symsubsup{e}{t}{i}(\symsubsup{z}{t}{i}),
\end{align}
with $a_k \in \{-0.8,-0.6,-0.4,-0.2,0,0.2,0.4,0.6,0.8\}$ and process noise covariance $\Sigma_k$ drawn from an $\mbox{IW}(3,0.5)$ prior. The time-series-specific features, shown in Fig.~\ref{fig:results1}(b), were sampled from a truncated IBP~\citep{GriffithsGhahramani:05} using $\alpha=10$
and then used to generate the observation sequences of Fig.~\ref{fig:results1}(a) (colored by the true mode sequences). Each row of the feature matrix corresponds to one of the five time series, and the columns represent the different autoregressive models with a white square indicating that a given time series uses that dynamical mode. Here, the columns are ordered so that the first feature corresponds to an autoregressive model defined by $a_1$, and the ninth feature corresponds to that of $a_9$.
\begin{figure}
[t!] \centering
\begin{tabular}
{c} \hspace{-0.1in}
\includegraphics[height = 2in]{figs/AR1_observations_model4}\\
(a)
\end{tabular}
\vspace{0.05in}
\begin{tabular}
{cc}
\includegraphics[height = 1.5in]{figs/trueF_model4_062609} &
\includegraphics[height = 1.5in]{figs/estF_model4_062609}\\
(b) & (c)
\end{tabular}
\caption[Synthetic data for 5 switching AR(1) time series, and associated true and BP-AR-HMM learned feature matrices.]{(a) Observation sequences for each of 5 switching AR(1) time series colored by true mode sequence, and offset for clarity. Images of the \mbox{(b)} true feature matrix of the five time series and (c) estimated feature matrix averaged over 10,000 MCMC samples taken from 100 trials every 10th sample. Each row corresponds to a different time series, and each column a different autoregressive model. White indicates active features. Although the true model is defined by only 9 possible dynamical modes, we show 20 columns in order to display the ``tail'' of the BP-AR-HMM estimated matrix resulting from samples that incorporated additional dynamical modes (events that have positive probability of occurring, as defined by the IBP prior.) The estimated feature matrices are produced from mode sequences mapped to the ground truth labels according to the minimum Hamming distance metric, and selecting modes with more than 2\% of the observations in a time series.} \label{fig:results1}
\end{figure}
The resulting feature matrix estimated over 10,000 MCMC samples is shown in Fig.~\ref{fig:results1}(c). Each of the 10,000 estimated feature matrices is produced from an MCMC sample of the mode sequences that are first mapped to the ground truth labels according to the minimum Hamming distance metric. We then only maintain inferred dynamical modes with more than 2\% of the time series's observations. Comparing to the true feature matrix, we see that our model is indeed able to discover most of the underlying latent structure of the time series despite the challenges caused by the fact that the autoregressive coefficients are close in value. The most commonly missed feature occurrence is the use of $a_4$ by the fifth time series. This fifth time series is the top-most displayed in Fig.~\ref{fig:results1}(a), and the dynamical mode defined by $a_4$ is shown in green. We see that this mode is used very infrequently, making it challenging to distinguish. Due to the nonparametric nature of the model, we also see a ``tail'' in the estimated matrix because of the (infrequent) incorporation of additional dynamical modes.
\subsection{Comparing the Feature-Based Model to Nonparametric Models with Identical State Spaces}
One might propose, as an alternative to the BP-AR-HMM, the use of an architecture based on the hierarchical Dirichlet process of~\citet{Teh:06}; specifically we could use the HDP-AR-HMMs of~\citet{Fox:IEEE11} tied together with a shared set of transition and dynamic parameters. For an HDP-AR-HMM truncated to $L$ possible dynamical modes, this model is specified as:
\begin{equation}
\begin{aligned}
\beta &\sim \mbox{Dir}(\gamma/L,\ldots,\gamma/L)\\
\pi_j \mid \beta &\sim \mbox{Dir}(\alpha\beta_1,\dots,\alpha\beta_{j-1},\alpha\beta_j + \kappa,\alpha\beta_{j+1},\dots,\alpha\beta_L)\\
\symsubsup{z}{t}{i} &\sim \pi_{\symsubsup{z}{t-1}{i}}, \quad
\symsubsupB{y}{t}{i} = \BF{A}_{\symsubsup{z}{t}{i}}\symsubsupB{\tilde{y}}{t}{i} + \symsubsupB{e}{t}{i}(\symsubsup{z}{t}{i}).
\end{aligned}
\end{equation}
Here, $\alpha$ and $\gamma$ are a set of concentration parameters that define the HDP and $\kappa$ is the sticky hyperparameter of the sticky HDP-HMM~\citep{Fox:AOAS11}; these hyperparameters are often given priors as well.
\subsubsection*{Segmentation Performance}
To demonstrate the difference between this HDP-AR-HMM and the BP-AR-HMM, we generated data for three switching AR(1) processes. The first two time series, with four times the data points of the third, switched between dynamical modes defined by $a_k \in \{-0.8, -0.4, 0.8\}$ and the third time series used $a_k \in \{-0.3, 0.8\}$. The results shown in Fig.~\ref{fig:results2} indicate that the multiple HDP-AR-HMM model, which assumes all time series share \emph{exactly} the same transition matrices and dynamic parameters, typically describes the third time series using $a_k \in \{-0.4, 0.8\}$ since this assignment better matches the parameters defined by the other (lengthy) time series. This common grouping of two distinct dynamical modes leads to the large median and 90th Hamming distance quantiles shown in Fig.~\ref{fig:results2}(b). The BP-AR-HMM, on the other hand, is better able to distinguish these dynamical modes (see Fig.~\ref{fig:results2}(c)) since the penalty in not sharing a behavior is only in the feature matrix; once a unique feature is chosen, it does not matter how the time series chooses to use it. Example segmentations representative of the median Hamming distance error are shown in Fig.~\ref{fig:results2}(d)-(e). These results illustrate that the IBP-based feature model emphasizes choosing behaviors rather than assuming all time series are performing minor variations of the same dynamics.
\begin{figure}[p!]
\centering
\begin{tabular}
{c}
\includegraphics[width = 0.85\columnwidth]{figs/model3obs}\\
(a)
\end{tabular}
\begin{tabular}
{cc} \hspace{-0.1in}
\includegraphics[height = 1.75in]{figs/HDPARHMM_IBPdata2_HammingDist_Seq3} &
\includegraphics[height = 1.75in]{figs/HammingDist_model3_1000trial}\vspace{-0.05in}\\
(b) & (c)\vspace{0.1in}\\
\includegraphics[width = 2.2in]{figs/HDPARHMM_IBPdata2_exampleSegmentation}\hspace{0.05in} &\hspace{0.05in}
\includegraphics[width = 2.2in]{figs/fullIBPHMM_IBPdata2_exampleSegmentation}\vspace{-0.05in}\\
(d) & (e)
\end{tabular}
\caption[Hamming distance quantiles comparing the segmentation performance of the HDP-AR-HMM to the BP-AR-HMM on a synthetic data example.]{(a) Observation sequences for each of 3 switching AR(1) time series colored by true mode sequence, and offset for clarity. The first and second sequences are four times as long as the third. (b)-(c) Focusing solely on the third time series, the median (solid blue) and $10^{th}$ and $90^{th}$ quantiles (dashed red) of Hamming distance between the true and estimated mode sequence over 1000 trials are displayed for the multiple HDP-AR-HMM model~\citep{Fox:IEEE11} and the BP-AR-HMM, respectively. (d)-(e) Examples of typical segmentations into behavior modes for the three time series at MCMC iteration 1000 for the two models. The top and bottom panels display the estimated and true sequences, respectively, and the color coding corresponds exactly to that of (a). For example, time series 3 switches between two modes colored by cyan and maroon.} \label{fig:results2}
\end{figure}
For the experiments above, we placed a $\mbox{Gamma}(1,1)$ prior on $\alpha$ and $\gamma$, and a $\mbox{Gamma}(100,1)$ prior on $\kappa$. The gamma proposals used $\sigma_{\gamma}^2=1$ and $\sigma_{\kappa}^2=100$ while the MNIW prior was given $M=0$, $K=0.1*I_d$, $n_0 = d+2$, and $S_0$ set to 0.75 times the empirical variance of the joint set of first-difference observations. At initialization, each time series was segmented into five contiguous blocks, with feature labels unique to that sequence.
\subsubsection*{Predictive Performance}
Using the same data-generating mechanism as used to generate the time series displayed in Fig.~\ref{fig:results2}(a), we generated a set of 100 held-out test datasets for Objects 1, 2, and 3. Each of the time series comprising the test datasets was of length 1000 (in contrast to the data of Fig.~\ref{fig:results2}(a) in which the time series of Object 3 was of length 500 and those of Objects 1 and 2 were of length 2000.) Based on a set of samples taken from 50 chains at MCMC iterations $[500~:~10:~1000]$ (i.e., a total of 2500 samples), we computed the log-likelihood of each of the 100 held-out datasets. That is, we added the time-series-specific log-likelihoods computed for each time series since the time series are conditionally independent given the model parameters. We performed this task for both the MCMC samples of the BP-AR-HMM and HDP-AR-HMM. The results are summarized in the histogram of Fig.~\ref{fig:hist_pred_perf}.
\begin{figure}[t!]
\centering
\includegraphics[width = 2.75in]{figs/pred_prob_model3_bin200}
\caption{Histogram of the predictive log-likelihood of 100 held-out data using the inferred parameters sampled every 10th iteration from MCMC iterations 500–1,000 from 50 independent chains for the BP-AR-HMM and HDP-AR-HMM run on the data of Fig.~\ref{fig:results2}(a).} \label{fig:hist_pred_perf}
\end{figure}
Since the BP-AR-HMM consistently identifies the unique dynamical mode of $a_k = -0.3$ used by Object 3 while the HDP-AR-HMM does not, we see from Fig.~\ref{fig:hist_pred_perf} that the mass of the BP-AR-HMM predictive log-likelihood is shifted positively by roughly 100 compared to that of the HDP-AR-HMM. In addition, we see that the histogram for the HDP-AR-HMM has a heavy tail, skewed towards lower log-likelihood, whereas the BP-AR-HMM does not.
Recall a couple of key differences between the BP-AR-HMM and HDP-AR-HMM. Both the HDP-AR-HMM and BP-AR-HMM define global libraries of infinitely many possible dynamic behaviors. However, the HDP-AR-HMM assumes that each of the time series selects the same finite subset of behaviors and transitions between them in exactly the same manner (i.e., the transition matrix is also global.) On the other hand, the BP-AR-HMM allows each time series to select differing subsets of behaviors \emph{and} differing transition probabilities. In the dataset examined here, the data-generating transition matrix between behaviors is the same for all time series, which matches the assumption of the HDP-AR-HMM. Second, two of the three time series share exactly the same dynamical modes, which is also close to the assumed HDP-AR-HMM formulation. The only aspect of the data that is better modeled apriori by the BP-AR-HMM is the unique dynamical mode of Object 3. However, there is not a large difference between this unique dynamic of $a_k = -0.3$ and the HDP-AR-HMM assumed $a_k = -0.4$. Regardless of the fact that the data are a close fit to the assumptions made by the HDP-AR-HMM, the improved predictive log-likelihood of the BP-AR-HMM illustrates the benefits of this more flexible framework.
\section{Motion Capture Experiments}
\label{sec:MoCap}
The linear dynamical system is a common model for describing simple human motion~\citep{Hsu:05}, and the switching linear dynamical system (SLDS) has been successfully applied to the problem of human motion synthesis, classification, and visual tracking~\citep{Pavlovic:99,Pavlovic:01}. Other approaches develop non-linear dynamical models using Gaussian processes~\citep{Wang:08} or based on a collection of binary latent features~\citep{Taylor:07}. However, there has been little effort in jointly segmenting and identifying common dynamic behaviors amongst a set of \emph{multiple} motion capture (MoCap) recordings of people performing various tasks. The BP-AR-HMM provides a natural way to handle this problem. One benefit of the proposed model, versus the standard SLDS, is that it does not rely on manually specifying the set of possible behaviors. As an illustrative example, we examined a set of six CMU MoCap exercise routines~\citep{CMUmocap}, three from Subject 13 and three from Subject 14. Each of these routines used some combination of the following motion categories: running in place, jumping jacks, arm circles, side twists, knee raises, squats, punching, up and down, two variants of toe touches, arch over, and a reach out stretch.
\begin{figure}[t!]
\centering
\includegraphics[width = 5.5in]{figs/MoCap2} \caption[Motion capture skeleton plots for BP-AR-HMM learned segmentations of six exercise routine videos.] {Each skeleton plot displays the trajectory of a learned contiguous segment of more than two seconds. To reduce the number of plots, we preprocessed the data to bridge segments separated by fewer than 300 msec. The boxes group segments categorized under the same behavior label, with the color indicating the true behavior label (allowing for analysis of split behaviors). Skeleton rendering done by modifications to Neil Lawrence's Matlab MoCap toolbox~\citep{LawrenceMoCap}.}\label{fig:MoCap}
\end{figure}
From the set of 62 position and joint angles, we selected the following set of 12 measurements deemed most informative for the gross motor behaviors we wish to capture: one body torso position, two waist angles, one neck angle, one set of right and left shoulder angles, the right and left elbow angles, one set of right and left hip angles, and one set of right and left ankle angles. The CMU MoCap data are recorded at a rate of at 120 frames per second, and as a preprocessing step we block-average and downsample the data using a window size of 12. We additionally scale each component of the observation vector so that the empirical variance on the concatenated set of first difference measurements is equal to one. Using these measurements, the prior distributions were set exactly as in the synthetic data experiments except the scale matrix, $S_0$, of the MNIW prior which was set to $5\cdot I_{12}$ (i.e., five times the empirical covariance of the preprocessed first-difference observations, and maintaining only the diagonal.) This setting allows more variability in the observed behaviors. We ran 25 chains of the sampler for 20,000 iterations and then examined the chain whose segmentation minimized an expected Hamming distance to the set of segmentations from all chains over iterations 15,000 to 20,000. This method of selecting a sample, first introduced in~\citet{Fox:AOAS11}, is outlined as follows. We first choose a large reference set $\mathcal{R}$ of state sequences produced by the MCMC sampler and a possibly smaller set of test sequences $\mathcal{T}$. Then, for each collection of state sequences $\BF{z}^{[n]}$ in the test set $\mathcal{T}$ (with $\BF{z}^{[n]}$ being the MCMC sample of $\BF{z}=\{z_{1:T}^{(i)}\}$ at iteration $n$), we compute the empirical mean Hamming distance between the test sequence and the sequences in the reference set $\mathcal{R}$; we denote this empirical mean by $\hat{H}_n$. We then choose the test sequence $\BF{z}^{[n^*]}$ that minimizes this expected Hamming distance. That is,
\begin{align*}
\BF{z}^{[n^*]} = \arg\min_{\BF{z}^{[n]} \in \mathcal{T}} \hat{H}_n.
\end{align*}
The empirical mean Hamming distance $\hat{H}_n$ is a \emph{label-invariant loss function} since it does not rely on labels remaining consistent across samples---we simply compute
\begin{align*}
\hat{H}_n = \frac{1}{|\mathcal{R}|} \sum_{\BF{z}^{[m]} \in \mathcal{R}}
\mbox{Hamm}(\BF{z}^{[n]},\BF{z}^{[m]}),
\end{align*}
where $\mbox{Hamm}(\BF{z}^{[n]},\BF{z}^{[m]})$ is the Hamming distance between sequences $\BF{z}^{[n]}$ and $\BF{z}^{[m]}$ after finding the optimal permutation of the labels in test sequence $\BF{z}^{[n]}$ to those in reference sequence $\BF{z}^{[m]}$. At a high level, this method for choosing state sequence samples aims to produce segmentations of the data that are \emph{typical} samples from the posterior. \citet{Jasra:05} provides an overview of some related techniques to address the label-switching issue.
The resulting MCMC sample is displayed in Fig.~\ref{fig:MoCap}. Each skeleton plot depicts the trajectory of a learned contiguous segment of more than two seconds, and boxes group segments categorized under the same behavior label by our algorithm. The color of the box indicates the true behavior label. From this plot we can infer that although some true behaviors are split into two or more categories by our algorithm, the BP-AR-HMM shows a clear ability to find common motions. Specifically, the BP-AR-HMM has successfully identified and grouped examples of jumping jacks (magenta), side twists (bright blue), arm circles (dark purple), squats (orange), and various motion behaviors that appeared in only one movie (bottom left four skeleton plots.) The split behaviors shown in green and yellow correspond to the true motion categories of knee raises and running, respectively, and the splits can be attributed to the two subjects performing the same motion in a distinct manner. For the knee raises, one subject performed the exercise while slightly twisting the upper in a counter-motion to the raised knee (top three examples) while the other subject had significant side-to-side upper body motion (middle three examples). For the running motion category, the splits also tended to correspond to varying upper body motion such as running with hands in or out of sync with knees. One example (bottom right) was the subject performing a lower-body run partially mixed with an upper-body jumping jack/arm flapping motion (an obviously confused test subject.) See Section~\ref{sec:chap5discussion} for further discussion of the BP-AR-HMM splitting phenomenon.
\begin{figure}[t!]
\centering
\includegraphics[width=.497
\textwidth]{figs/Hamm_GMM_HMM} \caption[Comparison of the BP-AR-HMM MoCap segmentation performance to HMM and Gaussian mixture model approaches.]{Hamming distance versus number of GMM clusters / HMM states on raw observations (blue/green) and first-difference observations (red/cyan), with the BP-AR-HMM segmentation (black, horizontal dashed) and true feature count (magenta, vertical dashed) shown for comparison. Results are for the most-likely of ten initializations of EM using an HMM Matlab toolbox~\citep{MurphyHMMtoolbox}.} \label{fig:GMM_Hamm}
\end{figure}
\begin{figure}
[t!] \centering
\includegraphics[width=.7
\textwidth]{figs/MoCap_F} \caption[Learned MoCap feature matrices from the BP-AR-HMM, HMM, and Gaussian mixture model approaches.]{Feature matrices associated with the true MoCap sequences (top-left), BP-AR-HMM estimated sequences over iterations 15,000 to 20,000 (top-right), and MAP assignment of the GMM (bottom-left) and HMM (bottom-right) using first-difference observations and 12 clusters/states.} \label{fig:GMM_F}
\end{figure}
We compare our MoCap performance to the Gaussian mixture model (GMM) method of~\citet{Barbic:04} using expectation maximization (EM) initialized with k-means.~\citet{Barbic:04}
also present an approach based on probabilistic principal component analysis (PCA), but this method focuses primarily on change-point detection rather than behavior clustering. As further comparisons, we consider a GMM on first-difference observations, and an HMM on both data sets. In Fig.~\ref{fig:GMM_Hamm}, we analyze the ability of the BP-AR-HMM, as compared to the defined GMMs and HMMs, in providing accurate labelings of the individual frames of the six movie clips~\footnote{The ability to accurately label the frames of a large set of movies is useful for tasks such as querying an extensive MoCap database (such as that of CMU) without relying on manual labeling of the movies.}. Specifically, we plot the Hamming distance between the true and estimated frame labels versus the number of GMM clusters and HMM states, using the most-likely of ten initializations of EM. We also plot the Hamming distance corresponding the BP-AR-HMM MCMC sample depicted in Fig.~\ref{fig:MoCap}, demonstrating that the BP-AR-HMM provides more accurate frame labels than any of these alternative approaches over a wide range of mixture model settings. The estimated feature matrices for the BP-AR-HMM and the GMM and HMM on first difference observations are shown in Fig.~\ref{fig:GMM_F}. The figure displays the matrix associated with the MAP label estimate in the case of the GMM and HMM, and an estimate based on MCMC samples from iterations 15,000 to 20,000 for the BP-AR-HMM. For the GMM and HMM, we consider the case when the number of Gaussian mixture components or the number of HMM states is set to the true number of behaviors, namely 12. By pooling all of the data, the GMM and HMM approaches assume that each time series exhibits the same structure; the results of this assumption can be seen in the strong bands of white implying sharing of behavior between the time series. The feature matrix estimated by the BP-AR-HMM, on the other hand, provides a much better match to the true matrix by allowing for sequence-specific variability. For example, this ability is indicated by the special structure of features in the upper right portion of the true feature matrix that is mostly captured in the BP-AR-HMM estimated feature matrix, but is not present in those of the GMM or HMM. We do, however, note a few BP-AR-HMM merged and split behaviors. Overall, we see that in addition to producing more accurate segmentations of the MoCap data, the BP-AR-HMM provides a superior ability to discover the shared feature structure.
\section{Discussion} \label{sec:chap5discussion}
We have presented a Bayesian nonparametric framework for discovering dynamical modes common to multiple time series. Our formulation reposes on the beta process, which provides a prior distribution on overlapping subsets of binary features. This prior allows both for commonality and time-series-specific variability in the use of dynamical modes. We additionally developed a novel exact sampling algorithm for non-conjugate IBP models. The utility of our BP-AR-HMM was demonstrated both on synthetic data, and on a set of MoCap sequences where we showed performance exceeding that of alternative methods. Although we focused on switching VAR processes, our approach could be equally well applied to HMMs, and to a wide range of other segmented dynamical systems models such as switching linear dynamic systems.
The idea proposed herein of a feature-based approach to relating multiple time series is not limited to nonparametric modeling. One could just as easily employ these ideas within a parametric model that pre-specifies the number of possible dynamic behaviors. We emphasize, however, that conditioned on the infinite feature vectors of our BP-AR-HMM, our model reduces to a collection of Markov switching processes on a \emph{finite} state space. The beta process simply allows for flexibility in the overall number of globally shared behaviors, and computationally we do not rely on any truncations of this infinite model.
One area of future work is to develop split-merge proposals to further improve mixing rates for high-dimensional data. Although the block initialization of the time series helps with the issue of splitting merged behaviors, it does not fully solve the problem and cannot be relied upon in datasets with more irregular switching patterns than the MoCap data we considered. Additionally, splitting a single true behavior into multiple estimated behaviors often occurred. The root of the splitting issue is two-fold. One is due to the mixing rate of the sampler. The second, unlike in the case of merging behaviors, is due to modeling issues. Our model assumes that the dynamic behavior parameters (i.e., the VAR process parameters) are identical between time series and do not change over time. This assumption can be problematic in grouping related dynamic behaviors, and might be addressed via hierarchical models of behaviors or by ideas similar to those of the \emph{dependent Dirchlet process}~\citep{MacEachern:99,Griffin:06} that allows for time-varying parameters.
Overall, the MoCap results appeared to be fairly robust to examples of only slightly dissimilar behaviors (e.g., squatting to different levels, twisting at different rates, etc.) However, in cases such as the running motion where only portions of the body moved in the same way while others did not, we tended to split the behavior group. This observation motivates examination of \emph{local partition processes}~\citep{Dunson:09,Dunson:09b} rather than \emph{global partition processes}. That is, our current model assumes that the grouping of observations into behavior categories occurs along all components of the observation vector rather than just a portion (e.g., lower body measurements.) Allowing for greater flexibility in the grouping of observation vectors becomes increasingly important in high dimensions.
\section*{Acknowledgments}
This work was supported in part by MURIs funded through AFOSR Grant FA9550-06-1-0324, ARO Grant W911NF-06-1-0076, and ONR Grant N00014-11-1-0688, and by AFOSR under Grant FA9559-08-1-0180 and Grant FA9550-10-1-0501. A preliminary version of this work (without detailed development or analysis) was first presented at a conference~\citep{Fox:NIPS09}.
|
\section{Introduction}
Due to their good isolation from the environment and to their
tunability, ultra-cold quantum gases are ideal candidates
to explore systems away from equilibrium~\cite{BlochZwerger2008}.
Cold atoms are
well suited to explore situations where the
Hamiltonian describing a system is slowly varied with time.
Understanding the physical implications of such slow quenches is of great
theoretical and practical importance to shed light on the coherent evolution
of quantum systems and to devise methods to prepare complex quantum phases.
Seminal works on the dynamics of classical
systems near a second order phase transition conducted by
Kibble~\cite{Kibble1976} and Zurek~\cite{Zurek1985} identified that the defect production
rate as a function of the ramp velocity is described by a scaling law when the system crosses a
critical point. However, despite many recent theoretical
advances~\cite{Dziarmaga2002, ClarkJaksch2004, Sengupta2004, ZurekZoller2005, Polkovnikov2005,
Cucchietti2007, Canovi2009, Dziarmaga2010, Pollmann2010, Zimmer2010, Zimmer2011,
Trefzger2011, Polkovnikov2011, BernierKollath2011, Dora2011, Kennett2011, Delande2009,
DziarmagaBishop2002, SchuetzholdFischer2006, Fischer2006, Moessner2010, RodriguezSantos2009, DeGrandi2010,
EcksteinKollar2009, MoeckelKehrein2010, EckardtHolthaus2005,EckardtLewenstein2011,PolettiKollath2011},
the response of strongly-correlated quantum gases to the slow
quench of a Hamiltonian parameter is still far from being fully understood. Meanwhile, on the
experimental side, considerable efforts have been devoted to understand the dynamics of
interacting bosonic atoms when the depth of the optical lattice is
varied~\cite{GreinerBloch2002, HungChin2010, Sherson2010, BakrGreiner2010, ChenDeMarco2011}
or when the slow quench of an effective parameter is
performed~\cite{ZenesiniArimondo2009,StruckSengstock2011}.
In relation to these experimental protocols, in Ref.~\onlinecite{HungChin2010, Rapp2010, NatuMueller2010, BernierKollath2011},
the presence of a parabolic trapping potential was found to significantly influence the dynamics.
Two dynamical regimes have been shown to exist when interacting atoms loaded into an optical lattice and confined to a trap
are subjected to a slow change of the interaction strength. For short ramp times, the evolution is dominated
by intrinsic local dynamics, which is also present in a homogeneous system, whereas, for
longer ramp times the density redistribution can play an important role.
In this work, we study the response of bosonic atoms to a linear change of the interaction
strength. As these atoms are confined to one-dimensional tubes and loaded into an optical
lattice running along the tubes main axis, the physics for a wide range of parameters
is well described by the one-dimensional
Bose-Hubbard model. Here our main objective is to understand the evolution, as a function of the
ramp time, of the local and non-local observables of the quantum gas. We focus on the crucial issues of the
formation and melting of Mott domains and on how the adiabatic limit is approached.
The article is structured as follows: In Sec.~\ref{sec:protocol}, we introduce the model and the time-dependent protocol.
Sections~\ref{sec:methods} and \ref{sec:continuity} detail the methods and the theoretical definitions. These two sections can
be skipped by readers more interested in the main phenomena. In Sec.~\ref{sec:SF_MI}, we turn
to the description of the evolution resulting from the increase of the interaction strength. In Sec.~\ref{sec:two_dyn}, we focus on the
occurrence of two dynamical regimes, the intrinsic dynamics and the dynamics induced by the
trapping potential, and explain their origin (Sec.~\ref{sec:pert}). Afterwards, we direct our attention to the formation
of ``Mott barriers'' which strongly block the equilibration process (Sec.~\ref{sec:form_MI}), the
energy transport (Sec.~\ref{sec:energy}) and the evolution of longer range correlations (Sec.~\ref{sec:corr}).
Then, in Sec.~\ref{sec:MI_SF}, we consider the opposite case of melting the Mott domains occurring when the
interaction strength is lowered and, in particular, we point out the long equilibration times for long
range correlation functions. For both situations, we characterize the time evolution considering various observables
such as the density, the compressibility, the energy, various particle
correlators, and the momentum distribution which is related to the interference patterns in time-of-flight
measurements. These results are supplemented by a detailed
analysis of the physical mechanisms responsible for the presence of the intrinsic and
global dynamical regimes. We further show that the different time-scales can be identified experimentally
from interference patterns. Our numerical results are obtained from the time-dependent density-matrix
renormalization group method (t-DMRG). We also compare these quasi-exact results to time-evolutions done within
the mean-field Gutzwiller method (Sec.~\ref{sec:gutz}) to identify the limitations of the
latter approach and pinpoint the qualitative physical insights it provides. The present
work extends substantially our previous results on the same setup~\cite{BernierKollath2011}.
\section{Model and conservation laws}
\label{sec:setup}
\subsection{Hamiltonian and time-dependent protocol}
\label{sec:protocol}
Ultracold bosons in optical lattices are, in a wide parameter regime, well described by the Bose-Hubbard Hamiltonian~\cite{FisherGrinstein1988,JakschZoller1998}:
\begin{equation*}
\mathcal{H} = -J \sum_l \left(\bhat^{\dag}_{l+1} \bhat_l + \text{h.c.}\right)
+ \frac{U(t)}{2} \sum_l \hat{n}_l(\hat{n}_l-1) -\sum_l \mu_l \hat{n}_l\,,
\end{equation*}
with $\bhat^{\dag}_l$ the operator creating a boson at site $l$ and $\hat{n}_l=\bhat^{\dag}_l \bhat_l$ the local density operator.
The total number of atoms is fixed to $N$.
The first term of the Hamiltonian corresponds to the kinetic energy of the atoms with the hopping amplitude $J$ and the second to the onsite interaction of strength $U$.
The site-dependent chemical potential $\mu_l$ accounts for an external confinement.
We consider here a one-dimensional geometry (tube) and use either a homogeneous ($\mu_l=0$) or a harmonic trapping potential of the form $\mu_l= -V_0(l-(L+1)/2)^2$, with $L$ the number of sites in the tube (open-boundary conditions) used in our simulations. We assume an experimentally realistic strength for the trapping potential of $V_0=0.006 J$ and particle number $N=24,48$. For these parameters the choice $L=64$ assures that edge effects are not important.
One non-trivial aspect of the model is that it is non-integrable~\cite{Kolovsky2004, Kollath2010} for non-zero $J$ and $U$. Further, at commensurate fillings, a quantum phase transition from a superfluid to a Mott-insulating state occurs (at $(U/J)_c \approx 3.4$ for unity filling in one-dimension~\cite{Kuhner2000, ZakrzewskiDelande2008}).
This phase transition is accompanied by the opening of a gap in the low-energy excitation spectrum, which strongly modifies the ground-state, thermodynamic and transport properties.
At incommensurate fillings a crossover between a superfluid and a Tonks-Girardeau, or hard-core boson, gas occurs in equilibrium.
This distinct behavior at commensurate and incommensurate fillings implies that in a trapped system, different states can coexist in spatially separated regions~\cite{BatrouniTroyer2002, KollathZwerger2003, Folling2006}.
For instance, for strong enough interaction, a Mott-insulating plateau with commensurate filling, surrounded by a superfluid region, emerges.
Regarding the time-dependent protocol, we consider a slow quench of the interaction strength $U(t)$ which can be achieved experimentally using a suitable Feshbach resonance~\cite{Inouye1998}. Different time-dependent protocols have been considered in previous works in homogeneous systems using several analytical or numerical approximation schemes~\cite{SchuetzholdFischer2006,DziarmagaBishop2002, Delande2009,Canovi2009,Cucchietti2007,RodriguezSantos2009,Dora2011,ClarkJaksch2004}.
For sake of simplicity and generality, the variation in time is chosen to be linear, starting from $U_i$ up to a final value $U_f$: $U(t) = U_i + \frac{t}{\tau}\delta U$ with $\tau$ the \emph{ramp time} and $\delta U = U_f-U_i$ the \emph{quench amplitude}.
The real-time evolution starts from the ground state corresponding to $U_i$.
The labels $i/f$ are used for the initial and final ground state values, respectively.
The limit $\tau \rightarrow 0$, i.e. the sudden quench limit, has been studied intensively in the Bose-Hubbard model using analytical~\cite{AltmanAuerbach2002, Polkovnikov2002, Fischer2006} and numerical methods~\cite{Kollath2007, LaeuchliKollath2008, Roux2009, Roux2010, Biroli2010, Sciolla2010, Sciolla2011}.
\subsection{Methods}
\label{sec:methods}
\subsubsection{t-DMRG}
Accurate ab-initio numerical simulations of the time evolution of the quantum gas are carried out using the t-DMRG technique~\cite{White1992, Vidal2004, WhiteFeiguin2004, DaleyVidal2004,Schollwoeck2005}. The time-evolution is implemented using the second order Trotter-Suzuki decomposition. The dimension of the effective space is a few hundred states and the time-step is adjusted with the ramp velocity. We introduce a cutoff value of $M=5$ or $6$ in the number of onsite bosons as higher boson occupancies are negligible in the situations considered here.
\subsubsection{Gutzwiller variational method}
In this section, we present how to determine the evolution of the system within the Gutzwiller
mean-field method \cite{RokhsarKotliar1991,JakschZoller2002}.
This approximation has been used before to describe the evolution during a slow change of the lattice depth in a higher dimensional trapped Bose-Hubbard system~\cite{Zakrzewski2005, NatuMueller2010}.
The Gutzwiller method is based on a variational ansatz of the many-body wave-function $|\Psi\rangle=\bigotimes_l \left[\sum_{n_l}c_{l,n_l}(t)|{n_l}\rangle\right]$ where $|{n_l}\rangle$ is the Fock state on site $l$ with $n_l$ particles and $c_{l,n_l}$ are the variational parameters.
The ground state for a given Hamiltonian is obtained by minimizing the total energy $E_{GW}$:
\begin{eqnarray*}
E_\text{GW} &=& -J \sum_l \left( \moy{\bhat_l}\!^* \moy{\bhat_{l+1}} + \text{c.c.} \right) \\
&&+ \frac{U(t)}{2} \sum_{l,n_l} n_l (n_l-1) |c_{l,n_l}|^2 - \sum_{l,n_l} \mu_l n_l|c_{l,n_l}|^2,
\end{eqnarray*}
where $\moy{\bhat_l} = \sum_{n_l} \sqrt{n_l+1} \; c^*_{l,n_l}c_{l,n_l+1}$ and $^*$ denotes complex conjugation.
The validity of the Gutzwiller method in evaluating static observables of one-dimensional systems has been studied, for example,
in Ref.~\onlinecite{Garcia-RipollvonDelft04}.
The Gutzwiller approach predicts, in one-dimension, a phase transition at $(U/J)_c = 2~(1+\sqrt{2})^2 \simeq 11.7$~\cite{RokhsarKotliar1991}.
The superfluid phase is signaled by a non-vanishing order parameter $\aver{\bhat_{l}}$ and, for a small interaction strength, the
properties of local quantities are reasonably well approximated.
In contrast, the Mott-insulating phase is characterized by a vanishing order parameter and vanishing local compressibility,
thus neglecting completely particle fluctuations which are present in the real Mott-insulating phase.
The time evolution for the coefficients $c_{l,n_l}(t)$ can be readily derived \cite{JakschZoller2002} from the Schr\"odinger equation. The equations are
\begin{eqnarray*}
i \hbar \partial_t c_{l,n_l}(t) &=& \left\{\frac{U(t)}{2} n_l (n_l-1) - \mu_l n_l\right\}c_{l,n_l}(t) \\
&& -J \sqrt{n_l+1} [\moy{\bhat_{l-1}}\!^* + \moy{\bhat_{l+1}}\!^*]\,c_{l,n_l+1}(t) \\
&& -J \sqrt{n_l} [\moy{\bhat_{l-1}} + \moy{\bhat_{l+1}}]\,c_{l,n_l-1}(t) \,.
\end{eqnarray*}
They are solved numerically by implementing a split-step method.
\subsection{Observables and definitions from continuity equations}
\label{sec:continuity}
\subsubsection{Correlations and interference pattern}
We define the one-body correlation function between sites $l$ and $m$
as
\begin{equation}
\label{eq:greens}
g_{l,m}= \frac{1}{2}\aver{\bhat_l^\dagger\bhat_{m} + {\rm h.c.}} \;.
\end{equation}
While $g$ is not easily accessible, time-of-flight techniques measure
an interference pattern related to the momentum distribution of the
correlated gas. Neglecting the Wannier-function envelope, the
interference pattern is given by
\begin{equation}
N(k) = \frac 1 L \sum_{l,m} e^{i(l-m)ka} \moy{\bhat_l^\dag \bhat_m}
\label{eq:nk}
\end{equation}
where $a$ is the lattice spacing.
For a superfluid state, $N(k)$ is expected to be strongly peaked
around zero momentum, whereas for a Mott-insulating state the
interference pattern should be rather flat.
\subsubsection{General expression for the continuity equation}
In this section, we want to determine, in the Schr\"odinger picture, the current operators corresponding to a given observable $O(t)=\elem{\psi(t)}{\hat{O}(t)}{\psi(t)}$, where $\hat{O}(t)$ can explicitly depend on time, using the associated continuity equation.
In the following, the shorthand notation $\moy{\cdots}$ stands for the expectation value $\elem{\psi(t)}{\cdots}{\psi(t)}$ at a given time.
The continuity equation takes, on general grounds, the following form:
\begin{align}
\hbar \partial_t O(t) \label{eq:continuity}
= -\text{div}\moy{\hat{J}^{O}} + \moy{\hat{S}^O}\,.
\end{align}
$\hat{J}^{O}$ is the current operator for which we have
\begin{subequations}
\begin{align}
-\text{div} \moy{\hat{J}^{O}} &=i \moy{\com{\Ham(t)}{\hat{O}(t)}}\\
&= -(\moy{\hat{J}^{O}_{l,l+1}} - \moy{\hat{J}^{O}_{l-1,l}}) \,. \label{eq:1dspez}
\end{align}
\end{subequations}
The second equality is the specialization to a one-dimensional lattice for an observable located around site $l$, with incoming and outgoing currents.
As we are interested in studying a one-dimensional system, \eqref{eq:1dspez} is used throughout.
The source operator $\hat{S}^O(t)=\hbar\partial_t\hat{O}(t)$ is non-zero only for an explicitly time-dependent operator.
Interestingly, integrating \eqref{eq:continuity} between times $0$ and $\tau$, taking the adiabatic limit $\tau\rightarrow\infty$ and
doing the change of variable $t\rightarrow U$ (in the source term integral) enables one to express the integrated contribution of currents only as a function of ground state expectation values
\begin{equation}
\int\limits_0^{\infty} \frac{dt}{\hbar} \,\text{div}\moy{\hat{J}^{O}}(t) = O_i - O_f + \int\limits_{U_i}^{U_f}\!\!dU\,\elem{\psi_0(U)}{\partial_U\hat{O}}{\psi_0(U)}\;,
\label{eq:adiabatic-O}
\end{equation}
where $\ket{\psi_0(U)}$ is the ground state corresponding to $U$.
The integral on the right-hand side is taken along the adiabatic path.
This remark is important from a numerical perspective because the right-hand side can be efficiently
computed via ground state techniques while the left-hand side would require time-dependent simulations
over very long times, which is not feasible.
As explained in the introduction, the main objective of this work is to characterize how particles and energy redistribute when the interaction strength, $U(t)$, is ramped up or down. Therefore, we introduce below the relevant quantities and the physically significant terms of their associated
continuity equations. A few commutators, useful in the derivation of these continuity equations, are provided in Appendix~\ref{app:commutators}.
\subsubsection{Local observables on sites and bonds}
As a first example, \eqref{eq:continuity} can be used to derive the particle current associated with the local density $\hat{n}_l$:
\begin{equation}
\hat{j}_{l,k}\equiv \hat{J}^{n_{l}}_{l,k} = iJ(\bhat^\dagger_{k}\bhat_{l}-\bhat^\dagger_{l}\bhat_{k}).
\label{eq:partcurrent}
\end{equation}
This current is defined between sites $l$ and $k$ and there is no source term associated with $\hat{n}_l$ as it is
not explicitly time-dependent.
As the particle current appears quite often in the rest of this article, from now on, it will be
denoted as $\hat{j}_{l,k}$. Finally, it is instructive to note that for a homogeneous and translationally invariant system,
the local density is constant at all time due to the conservation of the total number of particles.
In order to better understand the different time-scales involved during the evolution, it is also useful to consider separately
the evolution equation for the density fluctuations $\hat{n}_l^2$.
This quantity is essential to our comprehension of the Bose-Hubbard model and is related to the the local compressibility $\kappa_l = \aver{\hat{n}_l^2}-\aver{\hat{n}_l}^2$.
Using \eqref{eq:continuity} and after some algebra, we find that the evolution of $\hat{n}_l^2$ is controlled by a ``density-assisted'' or ``correlated'' current
\begin{equation}
\label{eq:evol_nn}
\hat{J}^{n^2_l}_{l,l+1}= \hat{n}_l\hat{j}_{l,l+1} + \hat{j}_{l,l+1}\hat{n}_l \;
\end{equation}
(note that $\hat{J}^{n^2_l}_{l-1,l} =\hat{n}_l\hat{j}_{l-1,l} + \hat{j}_{l-1,l}\hat{n}_l$).
The origin of this density-assisted current, mixing $\hat{j}$ and $\hat{n}^p$ operators, comes from the
evolution equation for the onsite occupancy probabilities discussed in Appendix~\ref{app:proba}.
In equilibrium, in our system, the average of~\eqref{eq:evol_nn} computed in
the ground state vanishes, as for the particle current operator.
The same strategy is used to get the current operators associated with observables living on bonds, such as the local kinetic energy operator $\hat{K}_{l,l+1}$ (or nearest-neighbor one-particle correlation) defined by
\begin{equation}
\hat{K}_{l,k} = -J(\bhat_k^{\dagger}\bhat_l + \bhat_l^{\dagger}\bhat_k)
\label{eq:kinetic}
\end{equation}
between site $l$ and $k$.
We find that the incoming current associated with $\hat{K}_{l,l+1}$ reads
\begin{subequations}
\label{eq:fullkineticcurrent}
\begin{align}
\hat{J}_{l-1,l}^{K_{l,l+1}} =& -\mu_{l} \hat{j}_{l,l+1} \label{eq:kin_mu} \\
& + J \hat{j}_{l-1,l+1} \label{eq:kin_corr2} \\
& + \frac{U(t)}{2} (\hat{n}_l\hat{j}_{l,l+1}+\hat{j}_{l,l+1}\hat{n}_l)\;, \label{eq:kin_densassisted}
\end{align}
\end{subequations}
showing the interplay of the correlated and usual particle currents.
It is worth noticing that for a homogeneous system, the evolution of local kinetic fluctuations is directly related to that of the density fluctuations since in this case
\begin{equation}
\label{eq:n2-kin-relation}
\partial_t \aver{\bhat_l^\dagger\bhat_{l+1}+\bhat_{l+1}^\dagger\bhat_l}= \frac{U(t)}{2J} \partial_t \aver{\hat{n}_l^2}\;.
\end{equation}
Thus, even in the homogeneous limit, the time dependence of $U(t)$ affects the evolution of the local kinetic term or nearest-neighbor correlations.
Eq.~\eqref{eq:n2-kin-relation} is also straightforwardly obtained from the evolution of the total energy discussed below.
Note, this equation is not valid for inhomogeneous gases where the balance of particle currents can be non-zero.
Similarly, the current operator associated with the particle current itself contains density-assisted hoppings, following the expression
\begin{subequations}
\label{eq:fullcurrentcurrent}
\begin{align}
\hat{J}_{l-1,l}^{j_{l,l+1}} =& \mu_{l} \hat{K}_{l,l+1} \label{eq:contr_mu_curr} \\
& + 2 J^2 \hat{n}_l \label{eq:contr_n_curr} \\
& + J \hat{K}_{l-1,l+1} \label{eq:contr_corr2_curr} \\
& - \frac{U(t)}{2} (\hat{n}_l\hat{K}_{l,l+1}+\hat{K}_{l,l+1}\hat{n}_l)\;. \label{eq:contr_densassisted_curr}
\end{align}
\end{subequations}
We also give for clarity the outgoing current operator: $
\hat{J}_{l+1,l+2}^{j_{l,l+1}} $ $
=\mu_{l+1}\hat{K}_{l,l+1}+2J^2\hat{n}_{l+1}+J\hat{K}_{l,l+2} $ $
-\frac{U(t)}{2}(\hat{n}_{l+1}\hat{K}_{l,l+1} $ $
+\hat{K}_{l,l+1}\hat{n}_{l+1}) $. It is worth mentioning that the
correlated current and hopping terms in \eqref{eq:fullkineticcurrent} and \eqref{eq:fullcurrentcurrent}
all come with the interaction strength as a prefactor and disappear
for a non-interacting gas. Their behavior is thus strongly affected by the presence of
interactions. Finally, as in the next section the time-derivative of the particle current
will be of great use to understand the mechanisms responsible for the evolution of the
density profile, we provide here its full expression:
\begin{subequations}
\label{eq:diffcurrent}
\begin{align}
\hbar \partial_t \moy{\hat{j}_{l,l+1}} =& (\mu_{l}-\mu_{l+1}) \moy{\hat{K}_{l,l+1}} \label{eq:contr_mu} \\
& + 2 J^2 (\moy{\hat{n}_l} - \moy{\hat{n}_{l+1}}) \label{eq:contr_n} \\
& + J (\moy{\hat{K}_{l-1,l+1}} - \moy{\hat{K}_{l,l+2}}) \label{eq:contr_corr2} \\
& -\frac{U(t)}{2} \moy{(\hat{n}_{l}-\hat{n}_{l+1})\hat{K}_{l,l+1} + \text{h.c.}}\;.\label{eq:contr_densassisted}
\end{align}
\end{subequations}
\subsubsection{Energy and heat}
\label{sec:energy}
We now turn to the transport of energy by first defining the bond-symmetric local energy operator as
\begin{equation}
\label{eq:local_e}
\hat{h}_l= \frac{1}{2}[\hat{K}_{l-1,l} + \hat{K}_{l,l+1}] + U(t)\hat{I}_l -\mu_l \hat{n}_l \; ,
\end{equation}
where $\hat{I}_l=\hat{n}_l(\hat{n}_l-1)/2$ is the operator related to the interaction energy.
In this case, we find that the energy current $\hat{J}^{h_l}_{l-1,l}$ is given by
\begin{subequations}
\label{eq:fullenergycurrent}
\label{eq:allenergy}
\begin{align}
\hat{J}^{h_{l}}_{l-1,l} =& -\frac{\left(\mu_{l-1}+\mu_l\right)}{2} \hat{j}_{l-1,l} \label{eq:energy_mu}\\
& -\frac{U(t)}{2} \hat{j}_{l-1,l} \label{eq:energy_j}\\
& -\frac{J}{2} (\hat{j}_{l-2,l} + \hat{j}_{l-1,l+1}) \label{eq:energy_corr2}\\
& +\frac{U(t)}{4} \left[ (\hat{n}_{l-1} + \hat{n}_l)\hat{j}_{l-1,l} + \hat{j}_{l-1,l}(\hat{n}_{l-1}+\hat{n}_l) \right]\,,\label{eq:energy_densass}
\end{align}
\end{subequations}
in which we naturally recover the particle and correlated currents appearing in \eqref{eq:partcurrent}, \eqref{eq:evol_nn} and \eqref{eq:fullkineticcurrent}.
In addition, since the energy operator is explicitly time-dependent and therefore not a conserved quantity during the protocol, we have the following source term
\begin{equation}
\label{eq:sourceE}
\hat{S}_l^{h_l} = \hbar \partial_t U(t) \hat{I}_l
\end{equation}
which shows the importance of the density fluctuations in the energy production.
In particular, the total energy $E(t)= \moy{\mathcal{H}(t)}$ satisfies the relation
\begin{equation}
\partial_t E(t) =\elem{\psi(t)}{\partial_t\Ham}{\psi(t)} = \left[ \partial_t U(t) \right] \sum_l \moy{\hat{I}_l}(t)\;,
\label{eq:total-energy}
\end{equation}
i.e., the energy put in the system is directly related to the evolution of the density fluctuations.
For an inhomogeneous system, there are two contributions to the local energy production as seen
from \eqref{eq:allenergy} and \eqref{eq:sourceE}: one from currents and correlated currents and one from the external driving of the system.
Summing up the total energy, the contribution from currents must vanish to fulfill \eqref{eq:total-energy}, but locally, one may have energy redistribution.
We can define the heat produced in the system as the energy of the atoms at the final time compared to that of the ground state for the final interaction strength
\begin{eqnarray}
Q(\tau) &=& E(\tau)-E_{0,f} \nonumber \\
&=& E_{0,i}-E_{0,f} + \frac{\delta{U}}{\tau} \int_{0}^{\tau} dt \sum_l \moy{\hat{I}_l}(t)\;,
\label{eq:heat}
\end{eqnarray}
with $E_{0,i/f}$ the ground state energies.
Note that $ \moy{\hat{I}_l} $ is accessible experimentally, which
makes it possible to measure the interesting $Q(\tau)$ dependence.
We can quickly check that this formula gives back the
correct results in the sudden quench and adiabatic quench limits.
In the sudden quench limit, $\ket{\psi(t)} = \ket{\psi_0(U_i)}$ which yields
$Q(0) = E_{0,i}-E_{0,f} + \delta U \sum_l \moy{\hat{I}_l}_{0,i}$.
This means that the heat only depends on ground state properties of the corresponding initial and final parameters.
In the adiabatic case, we have
$\ket{\psi(t)}=\ket{\psi_0(U(t))}$ along the adiabatic path so the
integral can be reexpressed as $\int_{U_i}^{U_f} dU \sum_l
\moy{\hat{I}_l}_0(U)$, with $\moy{\hat{I}_l}_0(U) = \elem{\psi_0(U)}{\hat{I}_l}{\psi_0(U)}$.
Using Feynman-Hellmann theorem over $U$, it is
clear that this integral cancels $E_{0,i}-E_{0,f}$ to make
$Q(\infty)=0$.
One can define a local excess energy $q_l$ as the difference in local
energies between the final energies and the ground state expectation
for the final parameters:
\begin{equation}
\label{eq:heatdef}
q_l(\tau) = \moy{\hat{h}_l}(\tau) - \moy{\hat{h}_l}_{0,f} \;.
\end{equation}
The local excess energy produced splits up into three different contributions
\begin{subequations}
\label{eq:allheat}
\begin{align}
\label{eq:heatdiff}
q_l(\tau) =& \moy{\hat{h}_l}_{0,i} - \moy{\hat{h}_l}_{0,f}\\
\label{eq:heatdiv}
& -\frac{1}{\hbar}\int_{0}^{\tau}dt \;\text{div}{\moy{\hat{J}^h}(t)} \\
\label{eq:heatfluct}
& + \frac{\delta U}{\tau}\int_{0}^{\tau} dt\; \moy{\hat{I}_l}(t)\;,
\end{align}
\end{subequations}
where the first term is simply the local ground state energies difference (independent of $\tau$),
the second term is the integrated contribution of energy currents, and the last term is the integrated
contribution due to the external operator. While $Q(\tau)$ is necessarily non-negative, $q_l(\tau)$ can
be negative or positive depending on the relative contributions of each term.
Finally, using \eqref{eq:adiabatic-O} with $\hat{O} = \hat{h}_l$ allows one to
calculate these quantities in the adiabatic limit:
\begin{equation}
\int\limits_0^{\infty} \frac{dt}{\hbar} \,\text{div}\moy{\hat{J}^{h_l}}(t)
= \moy{\hat{h}_l}_{0,i} - \moy{\hat{h}_l}_{0,f} + \int\limits_{U_i}^{U_f}\!\!dU\,\moy{\hat{I}_l}_0(U)\;,
\label{eq:adiabatic-energy}
\end{equation}
where the right-hand side can be computed accurately using numerical techniques.
With this set of equations in mind, we are now ready to identify the different driving
forces responsible for the system evolution when the interaction strength
is ramped up or down.
\section{Digging a Mott domain in a superfluid}
\label{sec:SF_MI}
\subsection{Evolution of local quantities from t-DMRG}
In this section, we consider a linear quench from $U_i=4J$ to $U_f=6J$.
$U_i$ is close to the homogeneous superfluid-Mott transition point and $U_f$ lies deeper in the Mott-insulating regime.
We compare two typical situations: (i) the number of particles is chosen low enough in order for the maximal
filling to remain below unity at all times ($N=24$); (ii) $N$ is sufficiently large so that, at $U_f$, the corresponding
ground state density profile has a Mott-insulating ``shell'' and a superfluid center ($N=48$).
We focus on different aspects of the dynamics: time-scales, role of insulating domains on particle
transport, energy production and transport, and their experimental signature.
\begin{figure}[t]
\centering
\includegraphics[width=\linewidth,clip=true]{fig1}
\caption{(color online). Slow quench from $U_i = 4J$ to $U_f = 6J$.
Evolution of local observables in the presence of a trap as a function of the ramp time
$\tau$, and compared with that of a homogeneous system (open symbols) having the same initial local density.
Observables are the density $n_l$, compressibility $\kappa_l$, occupancy probabilities $P_0$ and
$P_1$, neighboring correlation $g_{l,l+1}$ and the particle current $j_{l,l+1} = \moy{\hat{j}_{l,l+1}}$.
Subplots correspond to two different total number of particles $N=24$ and $48$, and two
different sites: $l=18$ and the central site $l= 32$ (cf.~Fig.~\ref{fig:trap_profiles_4_6} for the location of these sites).
\label{fig:local_vs_tau_4_6}}
\end{figure}
\subsubsection{Existence of two dynamical regimes}
\label{sec:two_dyn}
In Fig.~\ref{fig:local_vs_tau_4_6}, the final values ($t=\tau$) of most of the local observables introduced before
(density, local compressibility, local particle current, local correlation $g_{l,l+1}$) and also the first two
occupancy probabilities $P_0$ and $P_1$, are presented as a function of the ramp time $\tau$.
This figure clearly uncovers the existence of two dynamical behaviors.
First, we observe that for short ramp times, the densities at $l=32$ which lies in the center of the trap and $l=18$ which lies close to the forming Mott-insulating barrier are both approximately constant, following the evolution of the homogeneous system~\cite{Note1}.
In fact, variations (and oscillations) of both central and outer densities become significant only for longer ramp times, beyond $\hbar/J$.
In contrast, the evolution of both the occupancy probabilities and the compressibility occurs on a much faster time-scale: these observables vary rapidly at short $\tau$ and display less pronounced variations at larger $\tau$.
These two distinct behaviors reveal the presence of two dynamical regimes~\cite{NatuMueller2010, Rapp2010, BernierKollath2011}:
(i) the intrinsic dynamics, here occurring at short-times before the particle transport sets in (present in both the homogeneous and trapped systems);
(ii) a long-time behavior associated with particle transport and clearly due to the inhomogeneous structure of the density profiles.
Qualitatively, one can understand the origin of different time-scales from the continuity equations of Sec.~\ref{sec:continuity}.
For instance, the incoming and outgoing particle currents balance each other in a translational invariant configuration so that the density remains constant.
However, for the density fluctuations whose current operator~\eqref{eq:evol_nn} has correlated terms, no such balance
is achieved and consequently these quantities evolve with time.
In the case of an inhomogeneous gas, gradients of local quantities and chemical potentials inevitably give rise to particle currents, themselves sustaining the evolution of all local quantities.
Contrary to the intrinsic dynamics, we expect these effects to vanish with reducing the trap amplitude $V_0$.
Thus, their time-scale is distinct from the intrinsic one and is related to the external potential strength.
In order to quantify better these ideas, we now present arguments based on perturbative calculations.
\subsubsection{Insights from perturbative expansions}
\label{sec:pert}
In Ref.~\onlinecite{BernierKollath2011}, we observed that, for the quench
parameters typically considered, the
homogeneous dynamics of most local observables was well reproduced by
time-dependent perturbation theory, particularly in the small-$\tau$ regime. Working in the
\emph{initial Hamiltonian} eigenstates basis $\ket{\alpha}$, of energy
$E_{\alpha}$, the first-order expansion in $\delta U / \tau$ for a real,
symmetric and dimensionless observable $\hat{O}$ reads:
\begin{equation}
\label{eq:pert-O}
O(\tau,\delta U) = O_{00} - 2 \frac{\delta U}{\hbar\tau} \sum_{\alpha \neq 0} \frac{\omega_{\alpha}\tau-\sin(\omega_{\alpha}\tau)}{\omega_{\alpha}^2} O_{0\alpha}I_{\alpha 0}\;.
\end{equation}
The frequencies $\omega_{\alpha} = (E_{\alpha}-E_0)/\hbar$ are excitation
energies of level $\ket{\alpha}$ with respect to the ground-state $\ket{0}$.
$I_{\alpha\beta}=\sum_l\elem{\alpha}{\hat{I}_l}{\beta}$ are the matrix
elements of the interaction operator and
$O_{\alpha\beta}=\elem{\alpha}{\hat{O}}{\beta}$ those of the
observable. Such series can be well-behaved in the thermodynamic
limit even in the absence of a spectral gap, and this is what we
observe for our setup by looking at different system sizes. Taking
the $\tau \rightarrow 0$ limit~\cite{Note2},
the response of
the observable is typically quadratic
\begin{equation}
O(\tau,\delta U) \simeq O_{00}\left(1 \pm \frac{1}{2}f_{O}\tau^2\right)\,,
\end{equation}
where the $\pm$ sign depends on the observable. We have here introduce
the ``curvature''
\begin{equation}
f_O= \frac{2}{3}\frac{\delta U}{J} \frac{\tau_O^{-2}}{O_{00}}
\label{eq:fO}
\end{equation}
containing an intrinsic characteristic ramp time $\tau_O$
associated with the observable $O$ ($J$ is there for dimensionality
normalization):
\begin{equation}
\tau_O^{-2} = \frac{J}{\hbar} \Big\vert \sum_{\alpha \neq 0} \omega_{\alpha} O_{0\alpha}I_{\alpha 0}\Big\vert\;.
\label{eq:tauO}
\end{equation}
The curvature helps understand the departure from the initial value
$O_{00}$, as one can see, for example, in
Fig.~\ref{fig:local_vs_tau_4_6}. In particular, $f_O$ is linear with the
quench amplitude $\delta U$ (within this approximation).
For example, in the homogeneous gas limit, one can obtain an explicit expression
for the driving of the particle fluctuations $f_{n^2}$ by, in addition, resorting to
perturbation theory in $J/U_i$ (strong interaction limit).
We find that
\begin{equation}
f_{n^2}\approx \frac{32}{3}\frac{\delta U J^2}{\hbar^2 U_i}\;,
\label{eq:fn2}
\end{equation}
which is consistent with the change of the compressibility at short
time plotted in Fig.~\ref{fig:local_vs_tau_4_6}. The breakdown
of the quadratic behavior, which coincides with the onset of the relaxation,
is expected to happen on a time scale $\tau=|f_{n^2}|^{-1/2}$. We note that the
parameter $J/U_i$ in Fig.~\ref{fig:local_vs_tau_4_6} is not in the
regime where perturbation theory is expected to give a quantitative
description. Nevertheless, putting numerical values in
\eqref{eq:fn2}, one finds short relaxation times (below $\hbar/J$)
compatible with Fig.~\ref{fig:local_vs_tau_4_6}.
It is also worth mentioning that in the definition of the curvature
\eqref{eq:fO}, we were careful to separate what depends
on the quench protocol, the
parameter $\delta U$ and the prefactor $2/3$, from what is intrinsic
to the initial ground-state: $O_{00}$ and $\tau_O$.
Indeed, when the $U(t)$ function is of the general
type $\delta U f(t/\tau)$, the prefactor $2/3$ is replaced by $4~\int_0^1
dx(1-x)f(x)$. Hence, $\tau_O$ is
an intrinsic characteristic time of the initial state. We stress that the
quantities in \eqref{eq:fO} and \eqref{eq:tauO} are accessible by
ground state numerical techniques. Within this perturbative framework,
one can easily understand the two-regimes discussed above and also
derive, in the homogeneous case, relations between the characteristic
time-scales of various observables.
We first consider the characteristic time associated with the local density
operator. In the homogeneous case, the ground state is characterized by a
spatially uniform local density. Taking advantage of this symmetry, we
find that $\tau_n = \infty$. This result agrees with the fact that the
density remains constant for all times.
In contrast, for the local density fluctuations (or
compressibility) and local kinetic energy, $\tau_{n^2/g}$ are finite
even in a homogeneous system since the matrix elements in
\eqref{eq:tauO} do not vanish. Furthermore, the two time-scales are
actually related to each other. Since the Hamiltonian has only two
terms, we find that $\sum_l \elem{0}{\hat{g}_{l,l+1}}{\alpha} = (
U_i/J)\sum_l\elem{0}{\hat{n}_l^2}{\alpha}$ leading to $\tau_g =
\sqrt{\frac{J}{U_i}} \tau_{n^2}$.
This relation agrees with a dimensional
analysis of \eqref{eq:n2-kin-relation} and is also in qualitative
agreement with Fig.~\ref{fig:local_vs_tau_4_6}, where we find a
slightly slower relaxation for the kinetic term as compared to the
compressibility. In addition, these time-scales are themselves
related to the characteristic ramp time for the heat, $\tau_c$,
defined as $\tau_c^{-2}
= \frac{J}{12\hbar L}\sum_{\alpha} \omega_{\alpha}
\left\vert{I_{\alpha0}}\right\vert^2$~\cite{BernierKollath2011}.
Then, we find that $\tau_{n^2}
= \tau_c/\sqrt{24}$ (although the prefactor depends on the chosen
definition for $\tau_c$).
We now turn to the situation where a trapping confinement is present. In this case,
the translational symmetry is lost leading to a finite $\tau_n$.
Naturally, $\tau_{n^2/g}$ should also be affected by the presence of the trap,
but provided the latter is small enough, the corrections can be
negligible as illustrated by Fig.~\ref{fig:local_vs_tau_4_6}.
We expect that $\tau_n(V_0)$ diverges when the trap magnitude $V_0$
reaches zero. Consequently, by tuning $V_0$ to a low enough value, one
should in general be able to observe the intrinsic dynamics of the
system occurring below $\tau_n$. The behavior of the $\tau_n(V_0)$
function is an open issue, particularly because $V_0$ is not a
perturbation in experiments and realistic numerical calculations. If
we were to trust a naive first order perturbation argument for the
relatively unphysical situation of a gas in a box and perturbed by a
small $V_0$, one would expect a linear scaling of the matrix elements,
yielding the scaling $\tau_n \propto 1/\sqrt{V_0}$.
However, this scaling only serves as an illustration of the above statements.
Of course, $\tau_n$ also depends on $U_i/J$. Finally, we may argue that when
$V_0$ is too large, transport phenomena could eventually hide the intrinsic dynamics.
These results put on firmer grounds the existence of two
different dynamical regimes: one deeply connected to
inhomogeneities and controlled by $V_0$, and the intrinsic one
present in the homogeneous gas and much less sensitive to $V_0$.
\subsubsection{Profiles and ``Mott barriers''}
\label{sec:form_MI}
\begin{figure}[t]
\centering
\includegraphics[width=\linewidth,clip=true]{fig2}
\caption{(color online). Final local density and compressibility profiles after a slow quench
from $U_i=4J$ to $U_f=6J$ for different ramp times, $\tau$, and for the ground state ($\tau = \infty$)
at $U=6J$ in the trapped system. Left panels $N=24$, right panels $N=48$.
\label{fig:trap_profiles_4_6}}
\end{figure}
\begin{figure}[t]
\centering
\includegraphics[width=0.75\linewidth,clip=true]{fig3}
\caption{(color online). Time-evolution of the local compressibility $\kappa_l$ and current $j_{l,l+1} = \moy{\hat{j}_{l,l+1}}$
during a slow quench from $U_i=4J$ to $U_f=6J$ in a time $\tau = 7\hbar/J$ for $N=48$ in the trapped system.
As the ``Mott barriers'' are formed, the current in their vicinity weakens.
\label{fig:comp_vs_time}}
\end{figure}
We detail here the spatial evolution of local quantities.
In Fig.~\ref{fig:trap_profiles_4_6}, we present the final profiles for the density and compressibility as a function of ramp time.
At low filling ($N=24$), the shape of these final profiles is well understood if one resorts to the arguments presented above:
we see that for short ramp times the density profiles barely evolve while the compressibility changes considerably.
For longer $\tau$, the profiles approach smoothly the final ground state configuration.
For the larger filling $N=48$, the evolution is more involved.
A strong reduction of the compressibility occurs locally in regions of
filling close to unity already for short ramp times $\tau \approx \hbar/J$ while
the formation of pronounced Mott-insulating ``shells'' in the density profile only takes place at much longer ramp times, after $5\hbar/J$.
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip=true]{fig4}
\caption{$(a-d)$ Time-evolution of the different contributions to the time-derivative of the particle current:
Eq.~\eqref{eq:contr_mu} $(a)$, Eq.~\eqref{eq:contr_n} $(b)$, Eq.~\eqref{eq:contr_corr2} $(c)$
and Eq.~\eqref{eq:contr_densassisted} $(d)$. $(e)$ time-evolution of the time-derivative of the particle
current. $(f)$ time evolution of the particle current. For $(a)$ to $(e)$, the value at $t=0$ is subtracted.
The evolution parameters are the same as in Fig.~\ref{fig:comp_vs_time}.
\label{fig:contribution_current}}
\end{figure}
To understand better the complex dynamics at play in the presence of regions close to filling one,
we show in Fig.~\ref{fig:comp_vs_time} real-time snapshots of the compressibility and particle current
for the ramp time $\tau=7\hbar/J$.
The connection between these two quantities becomes evident at close inspection.
One first notices that, once again, the compressibility in regions away from unit filling evolves
quickly while the flow of atoms towards the system boundaries takes a much longer time to set in.
In addition, once the compressibility is sufficiently suppressed in the regions of filling one, the
current in these regions weakens, which slows down the density redistribution across the gas.
Even though the regions close to unit filling are small and are not real Mott-insulating plateaus, they
still reduce significantly the transport from the inner to the outer superfluid domains.
Consequently, the onset of low compressibility regions explains why systems above unity filling evolve
slowly when $U$ is increased.
From here on, we will refer to these regions as ``Mott barriers''.
To shed even more light on the build up and suppression of the particle current, we analyze the contribution of the different
terms appearing in \eqref{eq:diffcurrent} which make up the time-derivative of the particle current.
For each term, we plot in Fig.~\ref{fig:contribution_current}
the different contributions to $-\text{div}\moy{\hat{J}^j}$ for various times
in order to understand what drives the evolution of the particle current.
The first remarkable feature is Fig.~\ref{fig:contribution_current}(e) where the
time-derivative of the current changes sign near unity filling around $t\simeq 5\hbar/J$.
This inversion is a clear indication that the current is being suppressed by the formation of Mott barriers.
By considering each contribution separately using Fig.~\ref{fig:contribution_current}(a-d),
we observe that the main driving terms boosting the particle currents are the ones related to the
local kinetic energy \eqref{eq:contr_mu} and \eqref{eq:contr_corr2}, and the density assisted hoppings
term \eqref{eq:contr_densassisted} while the density gradients \eqref{eq:contr_n} become significant only
at the edges where the density varies rapidly.
The most striking phenomenon is due to the density assisted hoppings term \eqref{eq:contr_densassisted}.
We see on Fig.~\ref{fig:contribution_current}(d) that this term, which is non-zero only
in the presence of interactions, changes sign in the regions where Mott barriers are
forming thus drastically slowing down the equilibrating out-flow of atoms.
We finally stress again that the time-scales associated with the
contributions~\eqref{eq:diffcurrent} are essentially controlled by the steepness of profiles induced by the trapping potential $V_0$.
Within our choice of parameters they take longer times than the intrinsic evolution.
Experimentally, changing the confinement strength $V_0/J$ would affect both in time and magnitude
the creation of Mott barriers~\cite{Note3}.
\subsubsection{Energy transport and heat production}
We now turn our attention to the energy transport and heat production during a quench.
We present in Fig.~\ref{fig:trap_profiles_locale_4_6}(a, e) the final local energy profiles
for $N=24$ and $48$~\cite{Note4}.
This figure confirms our findings obtained from the analysis of the local density
and compressibility profiles: for $N=24$ the system approaches the adiabatic limit much faster than for $N=48$.
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip=true]{fig5}
\caption{(color online). Final local energy $(a,e)$ and local excess energy $(b,f)$ profiles after a
slow quench from $U_i=4J$ to $U_f=6J$ for different ramp times, $\tau$, and for the ground state ($\tau=\infty$)
at $U=6J$ for a trapped system. $(c)$ is the contribution
to $q_l$ due to the external operator
(see \eqref{eq:heatfluct}) while $(d)$ is the contribution from the energy currents (see \eqref{eq:heatdiv}).
The dashed line in $(b)$ and $(f)$ corresponds to $\moy{\hat{h}_l}_{0,i} - \moy{\hat{h}_l}_{0,f}$.
\label{fig:trap_profiles_locale_4_6}}
\end{figure}
For $N=48$, the final profile remains highly excited even for the longest ramp time considered ($\tau \approx 25 \hbar/J$).
The local excess energy production highlights a series of differences between the evolution of systems with
filling below and above one (see Fig.~\ref{fig:trap_profiles_locale_4_6}(b) and Fig.~\ref{fig:trap_profiles_locale_4_6}(f)).
We first notice that the local excess energy is smaller by nearly an order of magnitude for $N=24$ compared to $N=48$.
Furthermore, while for $\tau=25\hbar/J$ and $N=24$ the local excess energy is rather uniformly distributed
and close to zero, the $N=48$ result exhibits strong spatial fluctuations with $q_l$ large and negative at
the edges and large and positive at the center of the cloud. In fact, the local excess energy pattern
resulting from the quench is highly non-trivial even for the seemingly simplest situation where $N=24$, as
illustrated in Fig.~\ref{fig:trap_profiles_locale_4_6}(b, c, d).
At short ramp times (sudden quench limit), particles and energy currents are negligible so that the
term \eqref{eq:heatdiv} does not contribute, all the final excess energy being a balance between the ground
state energy difference \eqref{eq:heatdiff} and the density fluctuations average \eqref{eq:heatfluct}.
The latter is always positive and distributed rather uniformly in a Gaussian-like function whose maximum
decreases with $\tau$ (see Fig.~\ref{fig:trap_profiles_locale_4_6}(c)).
Hence, for short ramp times, the bulk retains most of the local excess energy while the
edges have negative $q_l$ due to the term \eqref{eq:heatdiff}.
For longer ramp times, energy currents set in with the effect of redistributing energy from the bulk to the
edges (see Fig.~\ref{fig:trap_profiles_locale_4_6}(b)).
Thus, these currents tend to strongly reduce both the spatial fluctuations and the total excess energy (heat) produced
by the quench. For intermediate times, either negative or positive $q_l$ at the edges and
in the bulk (see for instance the opposite distribution for $\tau = 10\hbar/J$ and $\tau = 15\hbar/J$ for $N=24$) can be found.
This effect arises as, for these parameters, the density profiles overshoot their final ground state configurations.
For $N=48$, the ``Mott barriers effect'' tends to freeze
the excess local energy pattern to the sudden quench typical distribution with negative $q_l$ at the edges
and positive in the bulk. Let us note that the freezing of the local excess energy pattern is strongly
related to the frozen density pattern.
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip=true]{fig6}
\caption{Time-evolution of the different contributions to the energy current:
$(a)$ Eq.~\eqref{eq:energy_mu}, $(b)$ Eq.~\eqref{eq:energy_j}, $(c)$
Eq.~\eqref{eq:energy_corr2} and
$(d)$ Eq.~\eqref{eq:energy_densass}. Time evolution of the full energy current $(e)$.
The evolution parameters are the same as in Fig.~\ref{fig:comp_vs_time}.
\label{fig:contr_energy}}
\end{figure}
Looking at the contributions to the evolution of the energy current in Fig.~\ref{fig:contr_energy},
we can identify the leading contribution driving the energy redistribution.
Comparing Fig.~\ref{fig:contribution_current}(f) and \ref{fig:contr_energy}(e), we see that the overall
evolution of the particle and energy current is very similar:
both currents set in at about the same time, and are suppressed in regions where Mott barriers form.
We also observe that the main contribution, determining the sign, to the energy current is from the density-assisted
particle current (see Fig.~\ref{fig:contr_energy}(d)).
However, this flow of energy towards the system edges is partially counterbalanced by two terms (Fig.~\ref{fig:contr_energy}(b)
and (c)) where the energy transport
occurs in the opposite direction to the particle current.
\begin{figure}[t]
\centering
\includegraphics[width=0.9\linewidth,clip=true]{fig7}
\caption{(color online). Total heat $Q(\tau)$ vs. inverse ramp time $\tau$ for both directions of the quenches for $N=24$ and $N=48$.
\label{fig:totheat}}
\end{figure}
Finally, one may wonder how the total heat produced $Q$ as a function of $1/\tau$ differs from the
homogeneous situation studied in Ref.~\onlinecite{BernierKollath2011}.
We show how the heat behaves as a function of the ramp time in Fig.~\ref{fig:totheat} for $N=24$ and $48$ for both the
quench from $U_i=4J$ to $U_f=6J$ and its reverse.
We first notice that
these curves are more complex than the one presented in Ref.~\onlinecite{BernierKollath2011} for a homogeneous system.
We also observe that the heat per atom produced in the case $N=24$ is always much lower than the one
for $N=48$. Our understanding of this phenomenon is that for lower filling the populated excited
states are less energetic as they are less likely to have doubly occupied sites.
We finally observe that, for fast ramps, the heat produced in the protocol with $U_i=4J$ and $U_f=6J$ is larger
than for the reverse protocol (while the opposite happens for slower ramps). We relate this to the fact that in the $U_i/J=4$
initial state a lot of particle fluctuations are present leading to a large interaction energy in the final state.
However, in order to fully understand the crossover to the inverse behavior at slow ramp times, the number of excitations
that are created and their final energies
would need to be identified, a task that we leave to future studies.
\subsubsection{Comparison with mean-field Gutzwiller method}
\label{sec:gutz}
\begin{figure}[t]
\centering
\includegraphics[width=0.9\columnwidth,clip=true]{fig8}
\caption{Time-evolution, within the Gutzwiller method, of a trapped one-dimensional Bose gas loaded
into an optical lattice during a quench from $U_i=6J$ to $U_f=15J$ for two different ramp times. Plotted
quantities are the local density $\langle \hat{n}_l \rangle$, the local compressibility $\kappa_l$,
the current $j_{l,l+1}$ and the superfluid order parameter $\langle\hat{b}_l \rangle$.
$N=54$, $L=84$, $V_t=0.006J$. Upper panels: $\tau=16\hbar/J$. Lower panels: $\tau=60\hbar/J$.
\label{fig:gutz}}
\end{figure}
Our aim here is to understand to what extent the mean-field Gutzwiller method can
describe the time-evolution of
a Bose gas loaded into a one-dimensional optical lattice and confined to a parabolic trap. With this objective
in mind, we study here a system made of $54$ atoms confined to a parabolic trap with $V_0=0.006J$ and loaded in an
optical lattice of $84$ sites, and consider slow quenches from $U_i = 6J$ to $U_f = 15J$ for two different
ramp times: $\tau=60\hbar/J$ and $\tau=16\hbar/J$.
These quenches begin on the superfluid side and the interaction strength is linearly increased up to a value
above the $n=1$ homogeneous superfluid-Mott-insulating transition, occurring at $U_c \approx 11.7J$ (using the Gutzwiller method).
At mean-field level, the ground state at $U = 6J$ is a superfluid with a
central density above one while the ground state at $U = 15J$ presents a broad Mott plateau.
Considering Fig.~\ref{fig:gutz}, we first notice that the Gutzwiller method captures well the presence of
two dynamical regimes. For both ramps, we see that the evolution of the local compressibility and superfluid order
parameter begins at $t = 0$ whereas the local density and the particle current remain fixed to their initial
values for a few $\hbar/J$. For a sufficiently fast quench, as shown in the upper panels of Fig.~\ref{fig:gutz},
we see that the superfluid order parameter, the compressibility and the current are strongly suppressed in a narrow
region around filling one. The local suppression of these three quantities around $ t = 6 \hbar/J$ signals the formation
of Mott barriers hindering the flow of atoms.
For $\tau=16J/\hbar$, these barriers are unstable and we notice the presence of oscillations reminiscent of the
ones arising when a strongly interacting phase is abruptly quenched to strong interactions \cite{Greiner2002b}.
By comparison, for sufficiently slow ramps, a stable Mott-insulating plateau forms at long times. On this plateau, the
condensate order parameter and the compressibility drop to zero. This total suppression of the density
fluctuations is an artifact of the mean-field method and also results in the absence of particle current on the plateau
as, within the Gutzwiller picture, the current factorizes into $j_{l-1,l}= 2J~\Im (\aver{\bhat_{l-1}}\!^* \aver{\bhat_l})$.
Finally, we also observe in the lower panels of Fig.~\ref{fig:gutz} that the quench triggers collective
breathing modes signaled by density oscillations
(along the time axis) in boundary regions~\cite{Note5}.
From this discussion of slow superfluid-Mott-insulating quenches within the Gutzwiller method, we conclude that this
approach captures some of the important out-of-equilibrium physical phenomena uncovered by t-DMRG, however as expected it cannot
provide an accurate quantitative picture.
\subsection{Evolution of non-local quantities}
\label{sec:corr}
\subsubsection{Real-space correlations}
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip=true]{fig9}
\caption{(color online). Value of the correlator $g_{l,l+d}$ with $l=L/2+1$ after a
slow quench from $U_i = 4J$ to $U_f = 6J$ in a trap for
various ramp times $\tau$ and the final ground state ($\tau=\infty$).
Left panel: $N=24$. Right panel: $N=48$.
\label{fig:trap_corr_long_range_MI}}
\end{figure}
Local and non-local correlations can propagate very differently during a quench.
To understand how correlations evolve during the slow quench of a global parameter,
we investigate here the evolution of single particle correlations.
Past studies on other systems have found that, after a slow parameter change, long-range
correlations take a long time to adjust~\cite{RodriguezSantos2009, ClarkJaksch2004}.
For example, for spin systems described by locally acting Hamiltonians, the propagation
of correlations during the slow change of a global parameter was found to be bound by a
``light-cone''. Outside of this light-cone, the so-called Lieb-Robinson bound, only
exponentially small changes to the correlations can be detected~\cite{LiebRobinson1972,BravyiVerstraete2006}.
We show in Fig.~\ref{fig:trap_corr_long_range_MI}
the value of the correlator $g_{L/2+1,L/2+1+d}$ at $t = \tau$ for different ramp times and two fillings.
The first striking result emerging from our study is that the evolution of this correlator is not
monotonic with the ramp time. We also find that in all cases the short distance correlator
responds quickly to the increase of the interaction strength, and that even for the fastest quenches
the final correlation function differs considerably from the initial ground state correlator.
Focusing on the left panel of Fig.~\ref{fig:trap_corr_long_range_MI}, we see that at low filling ($N=24$)
the short distance correlations reach their final ground state
values for almost all considered ramp times. In contrast, the longer range correlations
take much longer to reach their corresponding ground state values.
For example, for $\tau= 5 \hbar/J$, the long distance correlations have clearly not yet relaxed to their
final ground state values.
In the situation where regions with filling above one are present (right
panel of Fig.~\ref{fig:trap_corr_long_range_MI}), the evolution is even more involved. In this
case, the correlator at $t = \tau$ varies non-monotonically with distance and takes negative values
for intermediate ramp velocities. Even for the slowest ramps considered, the correlator
deviates considerably from its final ground state value at all distances. Finally, let us note that the
final ground state correlations present a dip at a distance corresponding to the location of
regions of filling one.
\subsubsection{Interference pattern for experiments}
\label{sec:interf}
Part of the complex dynamical behavior presented above can be observed experimentally in the
time-of-flight interference pattern $N(k)$ defined in \eqref{eq:nk}.
As shown on the left panel of Fig.~\ref{fig:trap_interf}, at low filling, the interference pattern present a peak at
$k = 0$ which changes in amplitude non-monotonically with the ramp time. This behavior
reflects the non-monotonic variation of long range correlations discussed in the previous section.
The final interference pattern is very different in the presence of regions close to filling one. In this case,
a peak at $k \neq 0$ develops at intermediate ramp times (see the right panel of Fig.~\ref{fig:trap_interf}).
This peak signals the strong out-of-equilibrium character of the state formed during the slow quench. However,
the absence of such a peak cannot be used to conclude that the system evolves adiabatically. Unfortunately, the
interference pattern is not as sensitive to out-of-equilibrium features as correlation functions are: $N(k)$
can be dominated by large ``in-equilibrium'' contributions coming from the short range correlators.
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip=true]{fig10}
\caption{(color online). Final value of the interference pattern $N(k)$ after a slow quench
from $U_i = 4J$ to $U_f = 6J$ in a trap for different ramp times
$\tau$ and for the final ground state ($\tau=\infty$).
Left panel: $N=24$. Right panel: $N=48$.
\label{fig:trap_interf}}
\end{figure}
\section{Melting of Mott-insulating regions}
\label{sec:MI_SF}
In this section, we consider a linear quench from $U_i=6J$ to $U_f=4J$. At $U_i$, the system presents a sizable
Mott-insulating ``shell'' and a superfluid center, while $U_f$ is close to the homogeneous superfluid-Mott-insulating transition
point. The ground state density and compressibility profiles at $U_f$ show none of the features associated
with the presence of Mott regions. Here again we focus on the different aspects of the dynamics:
time-scales, particle transport, energy production, and experimental signatures.
\subsection{Existence of two dynamical regimes}
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip]{fig11}
\caption{(color online). Slow quench from $U_i = 6J$ to $U_f = 4J$, $N=48$.
Evolution of local observables in the presence of a trap as a function of the ramp time $\tau$,
and compared with that of a homogeneous system (open symbols) having the same initial local density.
Observables are the density $n_l$, compressibility $\kappa_l$, occupancy probabilities $P_0$ and $P_1$,
neighboring correlation $g_{l,l+1}$ and particle current $j_{l,l+1} = \moy{\hat{j}_{l,l+1}}$.
\label{fig:n_vs_ts_melt}}
\end{figure}
When the interaction strength is lowered, the dynamics at play are also characterized by ``two dynamical regimes''.
However, as seen on Fig.~\ref{fig:n_vs_ts_melt}, in this case, the atoms are moving towards the center of
the system not towards the edges.
\subsection{Density, compressibility and energy profiles}
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip=true]{fig12}
\caption{(color online). Final profiles for the local density $(a)$, local compressibility $(b)$, local energy $(c)$ and
local excess energy $(d)$ after a slow quench from $U_i=6J$ to $U_f=4J$ for different ramp
times, $\tau$, and for the ground state ($\tau=\infty$)
at $U=4J$ for a trapped system. $(e)$ is the contribution to $q_l$ due to the external operator
(see \eqref{eq:heatfluct}) while $(f)$ is the contribution from the energy currents (see \eqref{eq:heatdiv}).
The dashed line in $(d)$ corresponds to $\moy{\hat{h}_l}_{0,i} - \moy{\hat{h}_l}_{0,f}$.
\label{fig:trap_profiles_6_4}}
\end{figure}
Considering the density and compressibility profiles for different ramp times shown in Fig.~\ref{fig:trap_profiles_6_4},
we notice that for ramp times of the order of $5\hbar/J$ the Mott-insulating regions are
almost fully melted and that the system is more
compressible. For example, local dips, initially present, have completely disappeared and only plateaus remain. For even
longer ramp times, the final density and compressibility profiles resemble closely the
$U_f$ ground state. Density redistribution occurs at a much faster pace when Mott-insulating regions are melted away than when they
are formed since in the former case Mott barriers are no longer effective. To conclude this comparison
between the two protocols,
it is interesting to note that, when the interaction strength is lowered, energy is transferred from the edges to the
center of the system as atoms pile up in the central region. The opposite occurs when the interaction is
increased.
\begin{figure}[t]
\centering
\includegraphics[width=\columnwidth,clip=true]{fig13}
\caption{(color online). Ramp from $U_i=6J$ to $U_f=4J$, $N=48$. Left panel: final value
for correlator $g_{l,l+d}$ with $l=L/2+1$ for different ramp times, $\tau$, and the
final ground state ($\tau=\infty$). Right panel: final value for the interference pattern $N(k)$ (see \eqref{eq:nk}) for
different ramp times and the final ground state ($\tau=\infty$).
\label{fig:trap_corr_long_range_6_4}}
\end{figure}
\subsection{Real space correlations and interference patterns}
Even though for long ramp times the density and compressibility profiles seem to evolve almost adiabatically,
the single particle correlator $g_{L/2+1,L/2+1+d}$ indicates that the system
is still far from equilibrium. As we can see on Fig.~\ref{fig:trap_corr_long_range_6_4}(a), this correlator is
negative at large distances and remains far from its ground state $U_f$ value even at long ramp times,
except for short distances. Therefore, to judge if the system has reached equilibrium by solely considering the
density and compressibility profiles is inadequate. Our results show unequivocally that the system is far from
having fully relaxed even at long ramp times. The non-equilibrium nature of the final state can be partially
probed by measuring experimentally the interference pattern (see~\eqref{eq:nk}).
On Fig.~\ref{fig:trap_corr_long_range_6_4} (right), we see that for intermediate ramp times,
the peak at $k = 0$
is shifted to higher momentum signaling the
non-equilibrium nature of the final state. However, as the interference pattern is a sum over all-distance
correlators and is dominated by short-distance values, it is difficult to distinguish the shifted peak
at long ramp times.
\section{Conclusion}
In this article we investigated the dynamics of the Mott-insulating regions of a bosonic gas
trapped and loaded into an optical lattice as the interaction strength is changed linearly with time.
We considered two situations: we first studied how Mott domains are formed by ramping up the
interaction strength from $U_i=4J$ to $U_f=6J$ and, in a separate set of simulations,
investigated how the domains melt when $U$ is ramped down. We conducted this study by examining
how the atomic density and compressibility profiles evolve, how particles and energy flow through the
system, how heat is produced and how single particle correlations propagate as a function of the ramp
time. For both situations we confirmed the existence of two dynamical regimes: an intrinsic regime
occurring at short times before particle transport sets in, and a long time behavior connected to
the system inhomogeneities and controlled by the strength of the underlying trapping potential.
We were able to establish the existence of these regimes on firmer grounds using various arguments
based on time-dependent perturbation theory. In a system with regions above unity filling, we found that a linear
increase of the interaction strength is accompanied by the formation of Mott insulating barriers
which hinder the flow of atoms from the center towards the edges. The emergence of these barriers is
evidenced by dips in the local compressibility and by the suppression of the particle current
in regions where the local density nears unity. We also established that, in these regions,
the change in sign of the particle current time-derivative is due to density assisted hopping,
a mechanism which only exists when $U$ is finite. The presence of theses barriers has multiple
consequences, among others, the system ``freezes'' into a highly excited configuration and
long range single particle correlations deviate strongly from their final ground state values,
even for the slowest ramp considered. This last feature could possibly be detected
experimentally from the gas interference pattern. By comparison, when the interaction strength is ramped down
the evolution is much less involved. For sufficiently long ramps, the density, compressibility and
local energy profiles approach the corresponding $U_f$ ground state configuration. However, even for the
slowest ramp considered, the final system is still far from being equilibrated as the associated one-body
correlator departs strongly from its final ground state value for all distances.
To conclude, we believe that this thorough investigation of the dynamics of a strongly interacting
bosonic gas will help experimentalists devise protocols to prepare complex quantum phases and provides
a new perspective to the understanding of the coherent evolution of quantum systems in the presence of
inhomogeneities.
\acknowledgments
We are grateful to R.~Citro, S.~Natu, E.~Orignac and A.~Rosch for fruitful discussions.
We acknowledge financial support from the Triangle de la Physique,
the Agence Nationale de la Recherche (under contract FAMOUS), the
SNSF (under division II), the DARPA-OLE program, and the Canadian
Institute for Advanced Research. Financial support for the computer
cluster on which some of the calculations were performed has been provided
by the Fondation Ernst et Lucie Schmidheiny.
|
\section*{Appendix}
For the readers convenience we very briefly sketch the elements of the argument relating
the MMP to the PCP. We consider the PCP over the two alphabets $\Sigma$
and $\Delta$, where $\Sigma$ is arbitrary and $\Delta=\{2,3\}$. Even though $\Delta$ is fixed, this version of the PCP
is still undecidable \cite{Halava}.
In order to relate this problem to a matrix problem, set $\Gamma\coloneqq\{1,2,3\}$ and
consider the map
$f:\Gamma^\ast \rightarrow {\mathbbm{N}}$ defined as
\begin{equation}
f(w) = \sum_{j=1}^{|w|} w_j 3^{|w|-j}
\end{equation}
for all nonempty words $w$ over $\Gamma$, where $|w|$ denotes the length of $w$.
$f(w)$ is the $3$-adic representation of $w$.
Now continue to define the
function $F: \Gamma^\ast \times \Gamma^\ast\rightarrow {\mathbbm{N}}^{3\times 3}$ as
\begin{equation}
F(u,v) = \left[
\begin{array}{ccc}
1 & 0 & 1\\
1 & 1 & 0\\
0 & 0 & 1\\
\end{array}
\right]
\left[
\begin{array}{ccc}
3^{|u|} & 0 & 0\\
0 & 3^{|v|} & 0\\
f(u) & f(v) & 1\\
\end{array}
\right]
\left[
\begin{array}{ccc}
1 & 0 & 1\\
1 & 1 & 0\\
0 & 0 & 1\\
\end{array}
\right]^{-1}.
\end{equation}
Let now $(h,g)$ be an instance of PCP, $h,g:\Sigma^\ast\rightarrow \Delta^\ast$.
For each such
instance, define the $3\times 3$-matrices
\begin{equation}
X_a= F(h(a),g(a)),\,
Y_a= F(h(a),1 g(a))
\end{equation}
for $a\in \Sigma$. Let $S$ be the matrix semigroup generated by $\{X_w,Y_w: w\in \Sigma\}$.
One then continues to consider matrix products
\begin{equation}
M= M_{w_1}\dots M_{w_n}\in S
\label{eqProduct}
\end{equation}
for a given word
$w=w_1\dots w_n$, where $M_{w_j}= X_{w_j}$ or $M_{w_j}= Y_{w_j}$.
The key step of the proof of Ref.~\cite{Halava}, deriving from the encoding of
Ref.~\cite{Paterson}, is to show that $M_{1,1}=0$ (denoting the upper left
element of the matrix) holds true
if and only if $w$ is a solution of the instance $(h,g)$. This shows that the problem to
decide whether the semigroup contains an element the upper left element of which
is zero is undecidable.
By adding the idempotent matrix
\begin{equation}
B= \left[
\begin{array}{ccc}
1 & 0 & 0\\
0 & 0 & 0\\
0 & 0 & 0
\end{array}
\right]
\end{equation}
as an additional generator to the set of matrices $\{X_w,Y_w: w\in \Sigma\}$,
it is then a simple step to reduce the MMP to the PCP.
In Ref.~\cite{Halava}, it is shown that we may choose $|\Sigma|=7$ and specific forms of the
product \eqref{eqProduct}, which gives a count of exactly $8$ matrix generators.
\end{document}
|
\section{Introduction}
Large-scale disturbances propagating through the solar corona have been first imaged by the
Extreme-Ultraviolet Imaging Telescope (EIT) onboard the Solar and Heliospheric Observatory (SOHO)
about 15 years ago \citep{thompson98}, and colloquially termed ``EIT waves''.
EIT waves have been initially interpreted as the coronal counterparts of Moreton waves,
in accordance with the fast-mode coronal MHD wave model developed by \cite{uchida68}.
However, due to statistical differences derived in the propagation velocities of Moreton waves, which are of the order of
1000 km~s$^{-1}$, and EIT waves, which lie mostly in the range 200--400 km~s$^{-1}$
\citep{klassen00,thompson09}, this interpretation was questioned and alternative models were developed.
Interpretations of EIT waves can be subdivided into wave versus non-wave models.
In the non-wave models, EIT waves are explained by the large-scale coronal restructuring
due to the erupting CME causing the observed emission enhancements either due to plasma compression, heating or localized energy release
\citep[e.g.][]{delanee99,chen02,attrill07}.
Coronal wave studies with the EIT instrument were limited by its 12~min observing cadence. This situation drastically improved by the launch of the
Solar Terrestrial Relations Observatory (STEREO) in October 2006, which has led to significant progress in the understanding of these intriguing events. The multi-point observing platforms of the twin STEREO satellites with their Extreme Ultraviolet Imagers \citep[EUVI;][]{howard08}
have made it possible to gain insight into the three-dimensional structure and evolution of EUV waves \citep{patsourakos09b,kienreich09,ma09,temmer11}, including events where the full three-dimensional coronal wave dome was observed \citep{veronig10}.
In addition, the high cadence (2.5 min), large field-of-view, and simultaneous observations from two vantage points facilitated the first detailed studies
on the wave dynamics and its relation to the associated CME evolution \citep{patsourakos09,kienreich09,temmer11}. Recent STEREO/EUVI findings include
observations of EIT waves, which undergo a significant deceleration during their propagation, accompanied by a decay of the perturbation amplitude and broadening of the wave front
\citep{veronig08,veronig10,long11,muhr11}. A set of four homologous EIT waves studied in \cite{kienreich11} revealed a distinct correlation between the wave speed and the magnetosonic Mach number.
These properties are consistent with the behavior of a (weak) fast-mode shock wave.
It is important to note that these recent observational improvements also led to significant advancement of MHD simulations of EIT waves \cite[e.g.][]{cohen09,downs11}.
For recent reviews see \cite{wills10}, \cite{gallagher10} and \cite{zhukov11}.
Since 2010, the Atmospheric Imaging Assembly \citep[AIA;][]{lemen11} onboard the Solar Dynamics Observatory (SDO) delivers ultra-high cadence (12~s) multiwavelength imaging of the solar atmosphere.
First studies of EIT waves with AIA revealed unexpected fine structures within the
diffuse wave front \citep{liu11}.
Due to the high sensitivity of AIA, also more EIT waves
with a three-dimensional dome are reported \citep[e.g.][]{kozarev11}.
What is still missing, however, is plasma diagnostics of EIT waves, i.e. density, temperature and plasma flows at the wave front. Such information cannot be obtained with EUV imagers but needs spectroscopic observations.
\cite{harra03} studied an EIT wave with the Coronal Diagnostics Spectrometer (CDS) onboard SOHO, without detection of
significant line-of-sight (LOS) velocities at the wave front (i.e. $v_{\rm LOS} \lesssim 10$~km~s$^{-1}$).
Two studies related to EIT waves were performed with the Extreme-Ultraviolet Imaging Spectrometer \citep[EIS;][]{culhane07} onboard Hinode \citep{asai08,chen10}
but the EIS observing mode was not suitable to study the plasma characteristics and motions at the EIT wave front.
In this letter, we present EIS plasma diagnostics
of the EIT wave of 2011 February 16 based on a unique data set obtained during our Hinode Observing Plan
HOP-180\footnote{http://www.isas.jaxa.jp/home/solar/hinode\underline{ }op/hop.php?hop=0180}, where we combined
high-cadence sit-and-stare EIS spectroscopy with high-cadence imaging by AIA.
EIS LOS plasma motions and line widths for the event under study have been presented in \cite{harra11}, whereas here we concentrate on the
relation between the different plasma parameters (including plasma densities and flows) at the wave front and their evolution.
\section{Data}
Hinode/EIS is a two-channel, normal-incidence EUV spectrometer observing in the wavelength ranges 170--210~{\AA} and 250--290~{\AA} \citep{culhane07}.
The high EIS spectral resolution allows Doppler velocity measurements of plasma flows better than $\pm$5~km~s$^{-1}$.
In order to study the plasma characteristics at propagating EIT wave fronts, we defined a dynamic EIS observing programme, which combines
high-cadence Hinode/EIS spectroscopy and SDO/AIA imaging. This programme was accepted as HOP-180
and realized during 2011 February 11--17. In HOP-180 we performed high-cadence EIS spectroscopic sit-and-stare observations (45~s exposure + 4~s readout time), placing the spectrometer slit (width: 1$''$, lengths: 512$''$) on the border of an active region, in order to follow the evolution of EIT waves along the EIS slit with a
spatial sampling of 1$''$ per pixel. We selected 11 EIS spectral lines over the temperature range log~$T = 4.7$ to 6.7,
including line pairs from the same ion for density diagnostics.
On 2011 February 16 EIS observed an EIT wave propagating along its slit (see Fig.~\ref{fig1}), on which we concentrate in this study.
Basic EIS photometric correction and correction of the orbital motion of the satellite were applied using the
eis\underline{ }prep.pro and eis\underline{ }wave\underline{ }corr.pro routines, before the spectral profiles
were fitted by a single Gaussian function with a linear background to obtain the
spectral intensities, integrated intensities, background intensities, Doppler shifts and spectral widths.
The zero reference of the Doppler shifts were calculated as the average value of the Doppler shifts
from quiet Sun regions. From the Fe\,{\sc xiii} 202/203~{\AA} line pair we derived the coronal electron densities from
the theoretical variation of the line ratio with density using the CHIANTI database version 6.0.1.\ \citep{dere97,dere09}.
Once the theoretical ratio was known, the final density maps were calculated using the EIS routine eis$\underline{~}$density.pro.
The EIT wave under study was also observed by SDO/AIA, which carries four (E)UV telescopes providing
full-Sun images in ten different wavelengths at a cadence as high as 12~s and spatial resolution of 1.5$''$ (with 0.6$''$ pixels). In this study, we use in particular the AIA 211~{\AA} (log~$T = 6.3$) and 193~{\AA} (log~$T = 6.1$) filters, in which the EIT wave signal was highest.
\section{Results}
Figure~1 shows a sequence of SDO/AIA 211~{\AA} direct and running ratio images together with the location of the Hinode/EIS
spectrometer slit. The EIT wave was launched in association with the M2 flare/CME event from AR~11158, and revealed a global propagation mainly towards
the Northern hemisphere. In particular, a distinct segment of the wave propagated along the EIS slit northward towards AR~11159 (see also movies no.~1 and~2).
Figure~\ref{fig2} shows stack plots of LOS velocities and intensities derived from the EIS Fe\,{\sc xiii}~202~{\AA} spectra.
During about 14:23 to 14:38~UT we observe a narrow lane of red-shifts indicating LOS velocities up to 20~km~s$^{-1}$ (Fig.~\ref{fig2}a) correlated with a bright lane
in the Fe~{\sc xiii} intensities (Fig.~\ref{fig2}b,c) with enhancements up to about 25\% above the pre-event level.
Co-aligned SDO/AIA images confirm that these signatures are due to the EIT wave front passing the EIS slit (see movie~1).
The peaks of the EIS intensity enhancements at the EIT wave front tend to occur delayed by one time step
($\approx$\,49~s) with regard to the EIS LOS velocity pulses, suggesting that the intensity change is a reaction to the downward push
of the coronal plasma below the wave front.
The brightest feature around 14:22~UT at $y \approx -255''$ is due to the associated M2 flare \citep{harra11}.
The EIT wave front visible as a narrow lane of red-shifts in the EIS LOS velocities is followed by blue-shifted pattern indicating relaxation
of the plasma behind the wavefront, with upward velocities $|v_{\rm LOS}| < 5$~km~s$^{-1}$ (Fig.~\ref{fig2}a).
Behind this lane of blue-shifts there is a broader lane of red-shifts indicating another propagating feature with downward-directed plasma motions and
concurrent intensity enhancements.
Since these LOS velocities are of the same order than that of the first lane of red-shifts, i.e. up to about 20~km~s$^{-1}$, they cannot be due to another swing
of the primary EIT wave but are rather due to a second disturbance moving behind.
In addition, there is a distinct dark lane observed in the EIS intensitygrams (Fig.~\ref{fig2}b), indicating propagation along the EIS slit with velocities of
$\approx$$120\pm 10$~km~s$^{-1}$, but with no reflection in the LOS velocities (Fig.~\ref{fig2}a,c). SDO/AIA images reveal that this signal is due to an ejecta moving northward along the slit (see accompanying movie no. 1).
Fig.~\ref{fig2}d shows stack plots of plasma densities derived from the EIS Fe\,{\sc xiii} 202/203~{\AA} line pair.
At the early phase of the EIT wave evolution, there is evidence of a lane of enhanced density
between about $y= -200''$ and $-100''$, roughly co-spatial with the propagating EIT wave.
However, careful comparison reveals that the enhanced densities are located about 2--3 time steps ($\sim$100--150~s) behind the wave front. At the positions of the density enhancements we observe decreased intensities and plasma upflows up to about $-50$~km~s$^{-1}$ (cf.~Fig.~\ref{fig2}a,b).
Thus, we interpret this enhanced density feature as being related to the eruption {\it behind} the EIT wave and not due to plasma
compression at the wave front itself. In the range $y=+120''$ to $+180''$ there is some indication of a
density depletion evolving with the EIT wave.
The distinct LOS velocity signal at the propagating EIT wave front is also well observed in the EIS Fe\,{\sc xii}~195~{\AA} spectral line (log $T = 6.11$)
but the LOS velocities are on average 5~km~s$^{-1}$ smaller than in the Fe\,{\sc xiii}~202~{\AA} line.
In Fe\,{\sc xvi}~262~{\AA} (log $T = 6.4$), a weak signature of the EIT wave can be observed.
There is no significant signal of the EIT wave in the He\,{\sc ii}~256~{\AA} line (log $T = 4.7$), indicating that the upper
chromosphere was basically unaffected by the passing EIT wavefront. This is consistent with the high-cadence H$\alpha$ observations by HASTA
(H$\alpha$ Solar Telescope for Argentinia), which show no signs of a Moreton wave.
We also note that in the hotter spectral lines we observed no clear signal related to the EIT wave except some signature in the Ca\,{\sc xvii}~192~{\AA} line (log~$T$ = 6.7), which we attribute to a blend by Fe\,{\sc xi}. Figure~\ref{fig3} shows cuts through the Fe\,{\sc xii} and Fe\,{\sc xiii} velocity stack plots, revealing the propagating LOS velocity pulse.
From these plots we derive the time, location and amplitude of the LOS velocity peaks used in the following to study the wave kinematics and the plasma characteristics
with EIS.
In addition, we calculated the EIT wave kinematics also from SDO/AIA 211~{\AA} running ratio images. The center derived from circular
fits to the earliest wave fronts ($x=462''\pm29''$, $y= -267''\pm 21''$) lies almost on
the EIS slit, and we thus calculated both the AIA and EIS wave kinematics as the
distance of the wave fronts along the spherical solar surface \citep[see][]{veronig06} to
the pixel located at the EIS slit at $x=440''$, $y=-267''$. The resulting SDO/AIA and Hinode/EIS kinematics are shown in Fig.~\ref{fig4}a, evidencing a strong deceleration of the EIT wave. The quadratic fit to the AIA measurements gives a start velocity $v_0 = 585\pm 56$~km~s$^{-1}$ and deceleration
$a = -675\pm 160$~m~s$^{-2}$; the mean velocity is $\bar{v} = 336\pm 15$~km~s$^{-1}$. These values are consistent with the results we obtain from calculating the EIT wave kinematics from the
positions of the peaks of the Hinode/EIS Fe\,{\sc xiii} LOS velocities, where we find $v_0 = 587\pm 50$~km~s$^{-1}$,
$a= -539\pm 48$~m~s$^{-2}$ and $\bar{v} = 371\pm 12$~km~s$^{-1}$.
Figure~\ref{fig4}b shows the evolution of the peak of the EIS Fe\,{\sc xiii} LOS velocity pulse at the wavefront, revealing an increase of the pulse during 14:24 UT up to 14:28 UT from
$\sim$3--5 to 15--20~km~s$^{-1}$, and a subsequent decay during the remaining wave evolution.
We also note that the EIT wave observed in AIA imagery revealed different characteristics in different propagation directions. The fastest propagation occurred into the NW quadrant, where we find
a start velocity $v_0 = 756\pm 56$~km~s$^{-1}$, deceleration $a=-489\pm 160$~m~s$^{-2}$ and mean velocity $\bar{v} = 576\pm 15$~km~s$^{-1}$.
If the EIT wave under study is due to a coronal fast-mode wave pushing the plasma downward and compressing it,
we expect a correlation between the amplitude of the velocity pulse and the change in plasma density and intensity at the wave front.
To test this hypothesis, we calculate the relative changes of these plasma parameters with respect to the
pre-event conditions by dividing each EIS density and intensity value by the corresponding pixel prior to the EIT wave.
The correlation plots for the EIS Fe\,{\sc xiii} parameters
are shown in the top panels of Fig.~\ref{fig5}, where we shifted the extracted EIS intensity values by one time step (49~s) with respect to the LOS velocities to account for their delayed response.
Indeed, for the relationship LOS velocity against intensity changes we obtain a positive correlation ($cc=0.45$) indicating that large downward velocities due to the passing EIT wave front are correlated with enhanced intensities, which are usually attributed to plasma compression \cite[e.g.][]{klassen00}. However, no such correlation is found between the LOS velocity and the relative changes in density.
This lack of correlation may be due to temperature enhancements as well as plasma compression contributing to the
intensity changes or due to uncertainties involved in the density derivation.
The maximum changes of intensities we observe at the EIT wave front are 25\% (except for one data point; see Fig.~\ref{fig5}a). Assuming that
the intensity enhancements at the wave front are solely due to enhanced plasma densities (and not due to changes in temperature), the maximum density increase expected is about 10\% ($I \propto n^2$ for $T=const$). This lies indeed within the noise level of our density estimates (cf.\ Fig.~\ref{fig2}d).
The bottom panels in Figure~\ref{fig5} show scatter plots of the LOS velocities against the absolute values of the EIS Fe\,{\sc xiii} intensities and densities at the EIT wave front,
revealing a distinct anti-correlation (intensities: $cc = -0.57$, densities: $cc = -0.49$).
The observed intensities and densities actually combine the line-of-sight integrated optically thin contributions of both the coronal ``background" plasma as well as the changes caused by the EIT wave.
The anti-correlation obtained for the absolute values may be related to the different states of ``background" corona through which the wave propagates.
Assuming conservation of the wave's energy flux $\rho v^2/2$, the LOS velocity caused by the passing wave front will be smaller when
propagating through a plasma of high magnetic and gas pressure such as in an active region (appearing as horizontal stripe
of high EUV intensity and enhanced density in Fig.~\ref{fig2}b,d) because the plasma is more inert
than in a quiet region where the magnetic and gas pressure are small.
This is a simplified picture, since also the wave properties themselves are expected to change during the propagation,
in terms of intensification and decay of the wave pulse, but it can qualitatively explain the observed anti-correlations.
\section{Discussion and Conclusions}
High-cadence EIS sit-and-stare spectroscopy combined with high-cadence AIA imaging allowed us to study
the plasma characteristics and evolution of a fast EIT wave that occurred on 2011 February 16. The fastest propagation was observed toward the NW quadrant from the source AR 11158 with a mean
velocity of 580~km~s$^{-1}$, whereas the propagation along the EIS slit toward AR~11158 in the Northern direction revealed a mean velocity of about 370~km~s$^{-1}$.
We observe downward plasma flows in coronal spectral lines formed at temperatures in the range 1.2--2.5~MK
co-aligned with the intensity enhancements at the propagating EIT wave front. On average the peaks of the EIS intensity
enhancements at the EIT wave front occur delayed by $\approx 1$~min with respect to the red-shifted velocity pulses,
suggesting that the intensity change is a reaction to the downward push of the coronal plasma below the wave front.
The peak LOS velocities of the downward plasma motions reach
values up to 20~km~s$^{-1}$ followed by upward velocities up to $-5$~km~s$^{-1}$ indicative of relaxation of the plasma behind the passing EIT wave front.
The downward plasma motions at the wave front reveal initial intensification and subsequent decay, correlated with the relative changes of the EIS spectral line intensities,
in line with the expectation for a fast-mode MHD wave.
However, no correlation is found between the relative density changes and the LOS velocities. This lack of correlation
is assumed to be due to the uncertainties of the density estimates
which are of the same order as the relative density enhancements expected from the observed wave front intensities (which are $\lesssim$25\%).
To settle this issue, spectroscopic observations of EIT waves of larger amplitudes are needed. For comparison, in the EIT waves
studied in \cite{muhr11}, relative intensity changes up to 70\% have been reported.
In addition to the decay of the LOS velocity pulse we also observe a significant deceleration of the EIT wave during its propagation ($a \approx -540$~m~s$^{-2}$).
Our findings are consistent with the interpretation that the EIT wave under study is a coronal fast-mode MHD wave,
pushing the plasma downward and compressing it at the coronal base though we note that definitive density diagnostics is still missing.
In contrast, non-wave models attributing the EIT wave to magnetic restructuring during the CME lift-off or to forced magnetic
reconnection ahead of the CME front do not predict a downward push of the lower corona.
In the He\,{\sc ii} line, no significant plasma motions at the EIT wave front are detected, consistent
with the fact that no H$\alpha$ Moreton wave was associated. This finding implies that in the EIT wave under study the
observed coronal wave pulse with LOS velocities $\lesssim$20~km~s$^{-1}$ was not strong enough to perturb the underlying ``dense'' chromosphere.
\acknowledgments AMV, PG, IWK, NM and MT gratefully acknowledge the Austrian Science Fund (FWF): P20867-N16 and V195-N16.
PG acknowledges support by VEGA grant 2/0064/09. We thank Dr.\ Wei Liu and the anonymous referee for insightful comments on the manuscript.
Hinode is a Japanese mission developed and launched by ISAS/JAXA, with NAOJ as domestic partner and
NASA and STFC (UK) as international partners. It is operated by these agencies in co-operation with ESA and NSC (Norway).
We thank the SDO/AIA team for designing and operating AIA.
CHIANTI is a collaborative project involving George Mason University,
the University of Michigan (USA) and the University of Cambridge (UK).
HASTA data are obtained at OAFA (El Leoncito, San Juan, Argentina) in the
framework of the German-Argentinean HASTA/MICA
Project, a collaboration of MPE, IAFE, OAFA and MPAe.
|
\section{Introduction}
Quantum fault tolerance \cite{G,ShorQFT,Preskill,AGP} is the framework which allows for accurate
implementation of quantum algorithms despite the inevitability of errors during the computation.
This is done by assuring that an error that occurs on one qubit cannot spread to multiple qubits. Application
of quantum error correction (QEC) then corrects the single qubit error \cite{book,ShorQEC,CSS}.
However, utilizing the entirety of the fault tolerant framework promises to be an expensive
proposition in terms of the number of qubits and implemented gates. Thus, it is worth exploring whether
it is possible to relax some of the strict rules required by the framework. One way to do this may
be by easing the construction requirements or simply not using Shor states as syndrome qubits
when encoding logical computational states and applying error correction. In this paper we study the
utilization of Shor states in the encoding of logical zero states and the application of error correction
for the [[7,1,3]] Steane code \cite{Steane} with the goal of limiting the number of required qubits and
implemented gates.
A fault tolerant method for encoding a logical computational state in the Steane code is to
apply fault tolerant error correction to any initial state of 7 qubits. This requires construction
of proper ancilla syndrome qubits such that each ancilla interacts with no more than one of the 7 data qubits.
For the Steane code there are a number of possible choices for these ancilla including Steane's \cite{SteaneAnc}
suggestion of using encoded ancilla blocks, and Knill's \cite{KnillAnc} method using encoded Bell states
and teleportation. In this work we have chosen to utilize four-qubit Shor states \cite{ShorQFT} for
ancilla as they require the least number of qubits and are thus most likely to be experimentally accessible.
Shor states are simply Greenberger-Horne-Zeilinger (GHZ) states with Hadamard gates applied to each qubit.
However, as the Shor states themselves are constructed in a noisy environment (here the nonequiprobable error environment), verification via parity checks is necessary to ensure accurate
construction. In this paper, we attempt to determine the number of Shor state verifications necessary to construct logical zero states or apply error correction with as high a fidelity
as possible. We then ask whether using Shor states with fewer verification steps (thus using fewer ancilla
qubits and requiring fewer gates) will provide sufficient accuracy to be used in the construction of
logical zero states or the application of error correction. Finally, we explore whether Shor states
are necessary at all in the construction of logical zeros and the application of error correction,
or whether sufficient accuracy may be obtained using single qubits for syndrome measurement.
The error model used in this paper is a non-equiprobable Pauli operator error model \cite{QCC} with non-correlated errors.
As in \cite{AP}, this model is a stochastic version of
a biased noise model that can be formulated in terms of Hamiltonians coupling the system to an
environment. In the model used here, however, the probabilities with which the different error
types take place is left arbitrary: the environment causes qubits to undergo a $\sigma_x^j$ error
with probability $p_x$, a $\sigma_y^j$ error with probability $p_y$, and a $\sigma_z^j$ error
with probability $p_z$, where $\sigma_i^j$, $i = x,y,z$ are the Pauli spin
operators on qubit $j$. We assume that only qubits taking part in a gate operation will be subject to error and
the error is modeled to occur after (perfect) gate implementation. Qubits not involved in a gate are
assumed to be perfectly stored. While this represents an idealization, it is partially justified in that
it is generally assumed that stored qubits are less likely to undergo error than those involved in gates
(see for example \cite{Svore}). In addition, in this paper accuracy measures are calculated only to
second order in the error probabilities $p_i$ thus the effect of ignoring storage errors is likely minimal.
Finally, we note that non-equiprobable errors occur in the initialization of qubits to the $|0\rangle$ state
and measurement (in the $z$ or $x$ bases) of all qubits.
This paper builds on the previous work of Ref.~\cite{YSW} (see also \cite{BHW}) in which the fault
tolerant method of encoding
logical zero states for the [[7,1,3]] code was compared to the gate sequence method of encoding to see
which method led to more accurately encoded zero states. Though the gate sequence method is not fault
tolerant (errors can propagate to multiple data qubits) it was found that the fidelity of the logical
zero states constructed in this way is comparable to the fidelity of the states constructed using the fault tolerant method.
Applying perfect error correction then revealed that the error probabilities were reduced to at least
second order for both methods (third order for the fault tolerant method), implying the correctability
of the errors and suggesting that either method can be used for practical quantum computation. Here we work
within the fault tolerant method in attempt to determine how best to construct Shor states for encoding
and error correction. In both papers, however, a major goal is to determine whether accurate enough protocols
can be implemented without invoking the full framework of quantum fault tolerance.
\section{Constructing Shor States}
\begin{figure}
\includegraphics[width=6.5cm]{Composite3b.eps}
\includegraphics[width=8.5cm]{Composite3a.eps}
\caption{Top: construction of a 4 qubit Shor state. \textsc{cnot} gates are represented by ($\bullet$) on the control qubit and ($\oplus$) on the target qubit connected by a vertical line. $H$ represents a Hadamard gate. The procedure entails constructing a GHZ state which is verified using ancilla qubits. Hadamard gates are applied to each qubit to complete Shor state construction.
Bottom: fault tolerant bit-flip and phase-flip syndrome measurements for the [[7,1,3]] code using Shor states (the Shor states pictured are assumed to have not had the Hadamard gates applied). To ensure fault tolerance, each Shor state ancilla qubit must interact with only one data qubit. The error syndrome is determined from the parity of the measurement outcomes of the Shor state ancilla qubits. To achieve fault tolerance each of the syndrome measurements is repeated twice.
Box: a useful equality which allows us to avoid implementing Hadamard gates by reversing the control and target of \textsc{cnot} gates. In our context, the \textsc{cnot} gates associated with the phase-flip syndrome measurements are reversed such that the ancilla qubits become the control and the data qubits become the target, as explained in the text. }
\label{ShorState}
\end{figure}
A construction method for the four-qubit Shor states needed for the [[7,1,3]] QEC code
is shown in Fig.~\ref{ShorState}. If the construction was done without error,
no verification steps would be needed and the Shor state (without the
final Hadamard gates as explained below) would be given by:
$|\psi_{Shor}\rangle = \frac{1}{\sqrt{2}}(|0000\rangle + |1111\rangle)$. However, actual
implementations of quantum computation will be done in a noisy environment and thus verifications
may be useful. We simulate construction of Shor states in the nonequiprobable error environment
including initialization and measurement errors with different verification strategies. We then
determine which of the strategies produce the highest quality Shor states based on the fidelity
of the constructed Shor states, the fidelity of logical zero states encoded fault tolerantly with
the different Shor states used as syndrome qubits, and the fidelity of a state after noisy error correction
when the different Shor states are used as syndrome qubits. The different strategies we
use are: no verifcation steps, one verification step, and different possible two verification steps.
The tenets of fault tolerance require that at
least one verification step be applied so as to lower the probability of error to second order.
\begingroup
\squeezetable
\begin{table*}
\caption{Relevant fidelity measures for Shor states and encoded logical zeros from different construction
methods: Shor state without
verification, Shor state with one verification, Shor state with two verifications, and the accuracy of
logical zero construction using single-qubit ancilla for syndrome measurements instead of Shor states.
The accuracy measures are the fidelity of the Shor state itself, the fidelity of the seven physical qubits
making up the logical zero state, the fidelity of the one qubit of information stored in the seven physical
qubits, and the fidelity after perfect error corrction has been applied to the constructed encoded zero states. }
\begin{tabular}{||c||c|c|c|c||}
\hline
& no verifications & 1 verification & 2 verifications & 1-Qubit ancilla \\\hline
\hline
Shor fidelity & $1-10p_x-11p_y-7p_z$ & $1-5p_x-6p_y-10p_z$ & $1-5p_x-6p_y-13p_z$ & \\\hline
7-Qubit fidelity & $1-85p_x-37p_y-12p_z$ & $1-55p_x-19p_y-12p_z$ & $1-55p_x-19p_y-12p_z$ & $1-49p_x-19p_y-12p_z$ \\\hline
1-Qubit fidelity & $1-25p_x-11p_y$ & $1-19p_x-7p_y$ & $1-19p_x-7p_y$ & $1-15p_x-7p_y$ \\\hline
after QEC & $1-92p_x^2-74p_xp_y-14p_y^2$ & $1$ & $1$ & $1-26p_x^2-6p_xp_y$ \\
\hline
\end{tabular}
\label{Tab1}
\end{table*}
\endgroup
To construct the Shor state we start with four qubits that we attempt to initialize to the state zero.
However, in this work, we assume that initialization itself is a noisy process subject to the same
error model as qubits involved in a gate. Thus, the actual state of each initialized qubit is
$\rho_i=(1-p_x-p_y)|0\rangle\langle 0|+(p_x+p_y)|1\rangle\langle 1|$. We then apply a Hadamard
gate, $H$, to the first qubit. The nonequiprobable error environment causes imperfections in the gate such
that the actual evolution of an attempted Hadamard on a single qubit $j$ in the state $\rho$ is:
\begin{equation}
\sum_{a=0,x,y,z}p_a\sigma_a^jH_j\rho H_j^{\dag}\sigma_a^j,
\end{equation}
where $\sigma_0^j$ is the identity matrix, $p_0 = 1-\sum_{\ell=x,y,z}p_\ell$,
and the terms $K_a^j = \sqrt{p_a}\sigma_a^jH_j$ can be regarded as Kraus operators for the
Hadamard evolution. The Hadamard is followed by a series of \textsc{cnot} gates. The attempted
performance of the \textsc{cnot} gate with control qubit $j$ and target qubit $k$, $\textsc{c}_j\textsc{not}_k$,
in the nonequiprobable error environment on any state $\rho$ actually implements:
\begin{equation}
\sum_{a,b=0,x,y,z}p_ap_b\sigma_a^j\sigma_b^k\textsc{c}_j\textsc{not}_k\rho \textsc{c}_j\textsc{not}_k^{\dag}\sigma_a^j\sigma_b^k,
\label{cnot}
\end{equation}
where terms $A_{a,b}^{j,k} = \sqrt{p_ap_b}\sigma_a^j\sigma_b^k\textsc{c}_j\textsc{not}_k$ can be regarded as the 16 Kraus
operators. Note that errors on the two qubits taking part in the CNOT gate are independent and not correlated. Shor state construction requires three \textsc{cnot} gates, shown in Fig.~\ref{ShorState}, and thus
the final Shor state is given by
\begin{eqnarray}
\rho_{Shor-err} &=& \sum_{a,b,c,d,e,f,g}^{0,x,y,z}A_{f,g}^{3,4}A_{d,e}^{2,3}A_{b,c}^{1,2}K_a^1\rho_i^{\otimes 4} \nonumber\\
&\times& (K_a^1)^\dag(A_{b,c}^{1,2})^\dag(A_{d,e}^{2,3})^\dag(A_{f,g}^{3,4})^\dag.
\end{eqnarray}
As explained above, applying the above described gate sequence does not guarantee that the resulting Shor states are suitable for fault tolerant quantum computation. During syndrome measurement, errors in the Shor state construction can
propagate into the data qubits. If only one Shor state qubit has been compromised
by error then only one data qubit will be compromised and the error can be subsequently
corrected. However, if multiple Shor state qubits are compromised, more than one data
qubit can be compromised and the computation will fail. Thus, we must test the Shor states
to ensure that multiple qubits have not been compromised by error.
This is done utilizing an ancilla qubit, initially in the
state $|0\rangle$, adjoined to the Shor state to measure the parity of random pairs of
qubits \cite{ShorQFT}. Should the test fail (the ancilla qubit measurement yields a
$|1\rangle$), the Shor state is immediately discarded. Of course, the ancilla qubit
initialization and the \textsc{cnot} gate implementations
for this parity check are themselves performed in the nonequiprobable
error environment and thus follow the dynamics described above.
We utilize an initial ancilla qubit to measure the parity of qubits 1 and 4.
Applying additional verification steps using additional ancilla may, if the \textsc{cnot}s themselves
are not too error prone, further ensure the lack of errors in the constructed Shor states.
A second ancilla can recheck the parity of the qubits checked with the first ancilla, or check the parity between other Shor state qubits. We have simulated every possible combination
for the second parity measurement and have found that this choice has little effect on any of our accuracy measures.
Our first accuracy measure for the Shor states constructed with different numbers of verifications
is the fidelity of the constructed Shor state as compared to a perfect Shor state,
$F = \langle\psi_{Shor}|\rho_{Shor-err}|\psi_{Shor}\rangle$. The fidelity results for
Shor states with zero, one, and two parity verifications are shown to first order in
error probability in the first row of Table \ref{Tab1}. Note that to first order the fidelity
for Shor states of two parity verifications is independent of which qubits are used for
the second verification.
Comparing the fidelity of the three Shor states we see that the Shor state with one verification
has a higher fidelity than the Shor state with no verifications unless $p_z$ is significantly
higher than $p_x$ and $p_y$. This demonstrates that it is usually advisable to perform a verification
step in order to suppress errors that occur during the Shor state construction. However, the fidelity of
the Shor state with two verification steps is always lower than that of the Shor state with one verification
step. A second verification step does not give enough benefit to outweigh additional errors that may occur
during the verification procedure.
\section{Encoding with Shor States}
The Shor state fidelity is a good measure of accuracy for the Shor state in and of itself.
However, our purpose for constructing Shor states is to use them to encode logical
zero states and implement fault tolerant error correction. It is possible that different errors in
the Shor state construction will have more or less of an effect on the accuracy with which
these protocols can be performed. Thus, another way to quantify the quality of the Shor states
is to simulate their utilization in the encoding of logical zero states
and in the performance of error correction and report on the accuracy with which these
protocols are implemented.
We first turn to the construction of logical zero states. To do this in a fault tolerant manner
we start with 7 qubits all noisily initialized to the state zero. Though this initialization is not perfect
we choose to not perform the first set of (bit-flip) syndrome measurements as their utility in
correcting an initialization error is outweighed by the noise inherent in applying the necessary
syndrome measurments.
Instead, we immediately measure the three phase flip syndromes (each one of the three twice) with Shor states as the syndrome qubits. To measure phase flip syndromes requires applying a Hadamard gate to each of the seven data qubits before and after the syndrome measurements. However, we can measure the syndrome without Hadamard gates if we reverse the roles of the control and target qubits for the \textsc{cnot} gates, and measure the Shor state qubits (noisily) in the $x$-basis, as explained in
\cite{Preskill} and shown in Fig.~\ref{ShorState}. For the case of encoding we analyze the scenario where all syndrome results are zero. Because
encoding is done `off-line' one can choose to utilize only the encoded states with this outcome.
Encoding in the nonequiprobable error environment using Shor states with different
numbers of applied verifications will result
in logical zero states with different degrees of accuracy. We can measure this accuracy in a number of ways.
The first way is simply to look at the fidelity of the seven qubit logical zero state. The accuracy of
this state gives an idea as to how well the entire encoding process was performed. Alternatively,
one may look at the fidelity of only the one qubit of encoded information. This is the only qubit
of information that is actually of importance and, if it is protected, the state of the rest of the system
is irrelevant. Measuring the fidelity of this one logical qubit is done by (noiselessly) decoding the
constructed logical zero state, tracing out all qubits but the first, and comparing the state
of the remaining qubit with the zero state on a single qubit. Both of these fidelity
measures have been calculated for logical zero states constructed using Shor states of zero, one,
and two verification parity checks, and are given in Table \ref{Tab1}.
Errors affecting the logical zero state may also be of varying degrees of severity. Applying perfect
error correction allows us to test the `correctability' of the types of
errors that occur during the encoding. If even perfect error correction cannot (to first order)
correct the errors in the logical zero state then the encoding method cannot be
used for practical implementations of quantum computation. We apply perfect error correction to
the states constructed using the Shor states with varying numbers of verifications
and calculate the fidelity measure of the output state. These fidelities are given in Table \ref{Tab1},
and corroborate our previous observations that applying one verifiction to Shor states is optimal.
Applying no verification steps to the Shor states leads to lower fidelities for the
logical zero states, and applying two verifications does not raise the fidelity. Perfect error
correction applied to logical zero states encoded using Shor states with one or two verifications
gives unit fidelity up to third order. However, perfect error correction applied to logical zero
states encoded using Shor states with no verifications, suppresses errors to second order implying
that these states may also be useable for practical quantum computation.
We compare the above cases of Shor state syndrome measurement with a logical zero encoding method
in which a single ancilla qubit is used for each syndrome measurement. This method does not meet the
standards of fault tolerance since an error on the single ancilla qubit, be it an initialization error or an
error in one of the syndrome measurement \textsc{cnot} gates, can spread to multiple data qubits. However,
using one ancilla qubit removes the need to construct Shor states thus lowering the
number of gates to be performed. The logical zero fidelity measures defined above are calculated for the single
qubit syndrome measurement construction method and are shown in Table \ref{Tab1}. Comparing these fidelity
measures to those calculated for Shor state based encoding, we find that using single qubit ancilla leads to higher
fidelity logical zero states. However, upon application of perfect
error correction the error probabilities are suppressed only to second order, unlike the logical zero states
constructed using Shor states for which the second order error probability terms are also suppressed.
\section{Quantum Error Correction with Shor States}
We now consider the accuracy with which the different Shor states can be used as syndrome ancilla qubits
for quantum error correction. The arbitrary single-qubit initial state we would like to protect is assumed
to have been perfectly encoded via the [[7,1,3]] gate encoding sequence:
$|\psi\rangle=\cos\alpha|0_L\rangle+e^{i\beta}\sin\alpha|1_L\rangle$, where
$|0_L\rangle$ and $|1_L\rangle$ represent the seven qubit logical $|0\rangle$ and $|1\rangle$ states
respectively. We assume the environment possibly causes an error such that, before error correction,
the system is in a mixed state of no error and all possible single qubit errors:
\begin{equation}
\rho_{err}=(1-7(p_x+p_y+p_z))|\psi\rangle\langle\psi|+\displaystyle\sum\limits_{i=1}^7\sum\limits_{a=x,y,z} p_a\sigma_a ^i |\psi\rangle\langle\psi|{\sigma_a^i}^{\dagger}.
\end{equation}
Because there are only single qubit errors in the system state, the error can be corrected by perfect
application of the [[7,1,3]] code.
\begingroup
\squeezetable
\begin{table*}
\caption{Fidelity measures for error correction applied to the state $\rho_{err}$ utilizing Shor states with different
numbers of verifications or a single ancilla qubit for syndrome measurement. In the Table $a = \cos[4\alpha]$ and
$b = \cos[2\beta]\sin[2\alpha]^2$. In this case the bit flip syndrome measurements were done first. }
\begin{tabular}{||c||c|c|c|c||}
\hline
& no verifications & 1 verification & 2 verifications & 1-Qubit ancilla \\\hline
\hline
7-Qubit fidelity & $1-85p_x-25p_y-7p_z$ & $1-55p_x-7p_y-7p_z$ & $1-55p_x-7p_y-7p_z$ & $1-49p_x-7p_y-7p_z$ \\\hline
& $1-(\frac{81}{4}+\frac{27}{4}a-\frac{27}{2}b)p_x$ & $1-(\frac{57}{4}+\frac{19}{4}a-\frac{19}{2}b)p_x$ & $1-(\frac{57}{4}+\frac{19}{4}a-\frac{19}{2}b)p_x$ & $1-(\frac{45}{4}+\frac{15}{4}a-\frac{15}{2}b)p_x$ \\
1-Qubit fidelity & $-(\frac{25}{4}+\frac{3}{4}a-\frac{5}{2}b)p_y$ & $-(\frac{13}{4}-\frac{1}{4}a-\frac{1}{2}b)p_y$ & $-(\frac{13}{4}-\frac{1}{4}a-\frac{1}{2}b)p_y$ & $-(\frac{13}{4}-\frac{1}{4}a-\frac{1}{2}b)p_y$ \\
& $-\frac{3}{2}(1-a)p_z$ & $-\frac{3}{2}(1-a)p_z$ & $-\frac{3}{2}(1-a)p_z$ & $-\frac{3}{2}(1-a)p_z$ \\\hline
\end{tabular}
\label{Tab2}
\end{table*}
\endgroup
To perform error correction in a fault tolerant manner, Shor states with at least one verification must
be used for syndrome measurements. We apply error correction to the state $\rho_{err}$ in the nonequiprobable
error environment by implementing
the three bit-flip syndrome measurements followed by three phase-flip syndrome measurements using Shor
states with different numbers of verifications as the syndrome qubits. Each syndrome
measurement is repeated twice to account for errors that may have occurred during the syndrome measurement
itself. We quantify the quality of the error correction via fidelity measures comparing the final state
after error correction to the pre-encoded arbitrary state.
\subsection{Syndrome Measruement Reveals No Error}
To read out the syndrome bit the four Shor state qubits are measured. If the results of the four measurements are of even parity the syndrome bit is a zero. If the results are of odd parity the syndrome bit is a one. We first look at the case where all qubit measurements are zero. Other even parity measurment results (say 0011 or 0101) give the same fidelity to first order. The fidelities of the seven data qubits and the one logical qubit state for this case are given in Table \ref{Tab2}.
Comparing the seven-qubit fidelities of the QEC procedure utilizing Shor states with different numbers of verifications, we first note that the fidelities
for Shor states with one and two verifications are identical up to second order terms. This fidelity is higher
than that attained by performing QEC using a Shor state with no verifications, again confirming that while
performing verification of the Shor state is important, there is no benefit gained from performing a second verification step. The fidelities exhibit little dependence on the initial state of the qubit, $\alpha$ and $\beta$ only appear in second order fidelity terms. Furthermore, regardless of the Shor state used, the $p_x$ error is dominant implying that
bit-flips are more harmful to the error correction procedure than phase flips.
Similar trends hold when comparing the single qubit fidelities except that all of the single
qubit fidelities depend strongly on the initial state. These fidelities are highest when $\alpha=0,\frac{\pi}{2}$, at which point the first and second order $p_z$ terms drop from the fidelity expression, and are lowest when $\alpha=\frac{\pi}{4}$.
Once again $p_x$ is the dominant error term. We note that the presence of first order terms
in the fidelity measures indicate that, in this case, noisy QEC cannot output a state with
no first-order error probability terms. Practical quantum error correction in this case is
thus reduced to minimizing the coefficients of these first order terms.
We compare the above QEC performance with that of error correction done without Shor states, instead using a single (noisily initialized) ancilla qubit for syndrome measurement. While this scheme certainly does not meet the criteria for fault tolerance, it does allow us to implement QEC with fewer qubits, and the lack of possible error from the construction of the Shor states may, and in fact does, yield an improved resulting fidelity. The fidelities for this case are shown in the last column of Table \ref{Tab2}.
\subsection{Syndrome Measruement Reveals Error}
Above we assumed that errors occur with low probability ($p_j \ll 1$) and thus the chances of measuring a bit-flip or phase-flip syndrome that is not 000 is extremely small. If, however, syndrome measurement does (twice in a row) signify an error a proper recovery operation must be performed. In such a case we find extremely low fidelities ($\approx .5$). The explanation for this is as follows (referencing Fig. 1, this should be compared to the fidelity results of \cite{BHW}): let us say that the syndrome 001 is measured (twice in a row) presumably indicating that the fourth qubit has undergone an error. Note, that the same syndrome would arise if an error had occurred to, say, qubit 7 during or after the $\textsc{c}_7\textsc{not}_{11}$ gate of the second syndrome bit. If this latter error had occurred the recovery operation would be applied to the wrong qubit, and, thus, the final state would have two errors: the error on qubit 7 which was not corrected and the error on qubit 4 due to the mistaken recovery operation. Because both the gates associated with error correction and the gates applied before error correction are implemented in the same error environment, there is no \emph{a priori} reason to think that this latter scenario is any less probable then the presumed error based on the syndrome measurement. Therefore, the final state of the system after the error correction procedure is a mixed state consisting of a corrected state and states with two errors (not to mention terms from errors that may have occurred during the final syndrome measurement, recovery operation, etc.) leading to an unacceptably low fidelity.
The above suggests a proper procedure to follow upon obtaining a non-zero syndrome measurement (even if the same syndrome is read out twice in a row). Rather then accepting the syndrome measurement and applying the requisite recovery operation, the syndrome measurement should be redone until the all zero readout is attained (twice in a row). Proceeding based on the non-zero readout will likely lead to a state with uncorrectable errors.
\subsection{Why Bit Flips?}
We noted above that $\sigma_x$ errors dominate the loss of fidelity. There are a couple of possibilities as to why this may be so. The first is because the bit-flip syndrome measurements were implemented first, and thus $\sigma_x$ errors that may occur during phase-flip syndrome measurements are not corrected.
A second possibility is that the use of (noisy) Shor states may cause the effect of $\sigma_x$ errors to be more
pronounced. In this section we clarify this issue by carrying out a series of simulations designed to isolate
the cause of increased sensitivity to $\sigma_x$ errors.
Our first step is to repeat the above error correction calculations implementing the phase-flip syndrome
measurements first. The fidelities of the resulting states are shown in Table \ref{Tab3}.
Let us first compare the cases where the syndrome measurement was done with a single ancilla qubit.
In this case the coefficients of the $p_x$ and $p_z$ terms in the seven qubit fidelity simply switch places
while the $p_y$ coefficient remains constant. Similarly in the one-qubit fidelity the $p_y$ coefficient
remains constant while the values of the $p_x$ and $p_z$ terms approximately trade values (modulo the
contribution of the initial state). This alone suggests that the dominance of the $p_x$ term
in the original simulations was simply because the bit-flip syndrome measurements were done first.
When the phase-flip syndrome measurements are done first $p_z$ replaces $p_x$.
However, when looking at the QEC simulations that utilize Shor states for syndrome measurements
we do not find the same trade-off. Instead, though the $p_z$ error coefficients
grow and (in most cases) become dominant, we find much less of a reduction of the
$p_x$ error coefficients. This suggests that there is something inherent in the use of the
(noisy) Shor states that leads to this type of error.
\begingroup
\squeezetable
\begin{table*}
\caption{Fidelity measures for error correction applied to the state $\rho_{err}$ utilizing Shor states with different
numbers of verifications or a single ancilla qubit for syndrome measurement. In the Table $a = \cos[4\alpha]$ and
$b = \cos[2\beta]\sin[2\alpha]^2$. In this case the phase flip syndrome measurements were done first. }
\label{fidtable}
\begin{tabular}{||c||c|c|c|c||}
\hline
& no verifications & 1 verification & 2 verifications & 1-Qubit ancilla \\\hline
\hline
7-Qubit fidelity & $1-61p_x-25p_y-55p_z$ & $1-31p_x-7p_y-55p_z$ & $1-31p_x-7p_y-55p_z$ & $1-7p_x-7p_y-49p_z$ \\\hline
& $1-(\frac{61}{4}-\frac{49}{4}a-\frac{3}{2}b)p_x$ & $1-(\frac{33}{4}-\frac{21}{4}a-\frac{21}{2}b)p_x$ & $1-(\frac{33}{4}-\frac{21}{4}a-\frac{3}{2}b)p_x$ & $1-(\frac{9}{4}+\frac{3}{4}a-\frac{3}{2}b)p_x$ \\
1-Qubit fidelity & $-(\frac{33}{4}-\frac{21}{4}a-\frac{1}{2}b)p_y$ & $-(\frac{13}{4}-\frac{1}{4}a-\frac{1}{2}b)p_y$ & $-(\frac{13}{4}-\frac{1}{4}a-\frac{1}{2}b)p_y$ & $(\frac{13}{4}-\frac{1}{4}a-\frac{1}{2}b)p_y$ \\
& $-\frac{27}{2}(1-a)p_z$ & $-\frac{27}{2}(1-a)p_z$ & $-\frac{27}{2}(1-a)p_z$ & $-\frac{25}{2}(1-a)p_z$ \\\hline
\end{tabular}
\label{Tab3}
\end{table*}
\endgroup
To further explore this point we perform two additional sets of QEC simulations. In the first,
we utilize perfect Shor states but allow errors (due to the nonequiprobable error environment) in
the error correction process (including syndrome measurement). In the second, we use Shor states
constructed in the nonequiprobable error environment (with one verification) but the error
correction itself (including syndrome measurements) is perfect. Both are done with bit-flip
syndrome measurements first and with phase-flip syndrome measurements first. When perfect
Shor states are used, but the error correction is noisy, we find that the dominant error
depends on which set of syndrome measurements is done first, if phase correction is done
first $\sigma_z$ errors dominate and vice-versa. The other error type is significantly diminished.
When noisy Shor states are used with perfectly implemented error correction we find that which
syndrome is done first makes little difference: $\sigma_x$ errors dominate and the fidelities do not
contain a first order term for $\sigma_z$ errors. The various fidelity measures are displayed in
Table \ref{Tab4}.
\begingroup
\squeezetable
\begin{table*}
\caption{Fidelity measures for quantum error correction applied to the state $\rho_{err}$ with perfect Shor
states and noisy error correction, and noisy Shor states with perfect error correction. Both cases were done
with the $\sigma_x$ syndrome measurements first and the $\sigma_z$ syndrome measurements first. In the Table
$a = \cos[4\alpha]$ and $c = \cos[2\beta]$.}
\begin{tabular}{||c||c||c|c||}
\hline
& & bit-flip first & phase-flip first \\\hline
\hline
& 7-Qubit fidelity & $1-31p_x-7p_y-7p_z$ & $1-7p_x-7p_y-55p_z$ \\\cline{2-4}
Noisy QEC & & $1-(\frac{33}{4}+\frac{11}{4}(a-c+ac)p_x$ & $1-(\frac{9}{4}+\frac{3}{4}a-\frac{3}{2}b)p_x $ \\
Perfect Shor States & 1-Qubit fidelity & $-(\frac{13}{4}-\frac{1}{4}(a+c-ac)p_y$ & $-(\frac{13}{4}-\frac{1}{4}a-\frac{1}{2}b)p_y$ \\
& & $-\frac{3}{2}(1-a)p_z$ & $-\frac{27}{2}(1-a)p_z$ \\\hline
Perfect QEC & 7-Qubit fidelity & $1-24p_x$ & $1-24p_x$ \\\cline{2-4}
Noisy Shor States & 1-Qubit fidelity & $1-(6+2a-4b)p_x$ & $1-(6+6a)p_x $ \\\hline
\end{tabular}
\label{Tab4}
\end{table*}
\endgroup
Taken together these simulations imply that when noisy Shor states are utilized for syndrome measurement
in the Steane code there is a significant bias towards bit-flip errors. A possible solution is to
concatenate into a three-qubit bit-flip QEC code for another level of error correction. This could
significantly reduce the sensitivity to bit-flip errors without the resource cost of concatenation into
another level of the seven-qubit Steane code.
\section{Conclusion}
In conclusion, we have calculated quality metrics for different Shor states used as syndrome
measurement ancilla qubits for the [[7,1,3]] CSS QEC code operating in a nonequiprobable error
environment. The results suggest that while a Shor state constructed in this error environment
with one parity check verification is optimal for suppressing errors in the construction of
logical zero states, Shor states with no checks will also suppress error probability terms in the fidelity to second order. In addition, encoding applied without Shor states, instead using single qubit ancilla for syndrome measurement, leads to logical zero states with
higher fidelity but errors that are less correctable as identified by fidelity after perfect
error correction.
For error correction applied in a nonequiprobable error environment using the seven qubit Steane code,
our simulations show that not using Shor states leads to a corrected state with higher fidelity than using Shor states. In addition, we noted that bit-flip errors are dominant whether
Shor states are used or not. We first suggested that this was due to the fact that the bit-flip syndrome measurements were
done first, meaning that uncorrected bit-flips may accumulate during phase-flip syndrome measurements.
Simulations switching the order of the syndrome measurements demonstrated that this is correct when
using single qubit ancillae for syndrome measurement, but does not completely explain the results of
simulations using Shor
states. Further simulations indicated an inherent sensitivity towards bit-flip errors when Shor states
are used. We suggested that this could be overcome by concatenating with a three-qubit QEC code
that protects against bit-flip errors. Finally, we suggested that when a non-zero syndrome is detected implementing the prescribed recovery operation will lead to a state of unacceptably low fidelity. Rather the syndrome measurement should be repeated until a zero syndrome readout is attained.
The authors would like to thank G. Gilbert for constructive comments. This research is supported under MITRE Innovation Program Grant 07MSR205.
|
\section{Introduction}
The goal of the present work is to give an explicit construction of a representation of the positive quantum group $GL_q^+(N,\R)$ and its modular double $GL_{q\til[q]}^+(N,\R)$ by positive essentially self-adjoint operators acting on certain Hilbert space $\cH$. This is done by finding a quantum analogue of the Gauss-Lusztig decomposition for $GL_q(N)$.
Let $G$ be a semi-simple group of simply-laced type, $T$ its $\R$-split maximal torus of rank $r$, and $U^{\pm}$ its maximal unipotent subgroup with $\dim U^+=m$. The Gauss decomposition of the max cell of $G$ is given by
\Eq{ G=U^- T U^+.}
In type $A_r$ $(N=r+1)$, this amounts to the decomposition into lower triangular, diagonal and upper triangular matrices.
On the other hand, given a \emph{totally positive} matrix $G_{>0}$, where all entries of the matrix and the determinants of its minors are strictly positive, it can be decomposed as
\Eq{ G_{>0}=U_{>0}^- T_{>0} U_{>0}^+,}
where all the entries and determinant of the minors of $U^{\pm}$ and $T$ are strictly positive if they are not identically zero. Lusztig in \cite{Lu} discovered a remarkable parametrization of $G_{>0}$ using a decomposition of the maximal Weyl group element $w_0\in W$. Let $w_0=s_{i_1}...s_{i_m}$ be a reduced expression for $w_0$, then there is an isomorphism between $\R_{>0}^m\to U_{>0}^+$ given by
\Eq{(a_1,a_2,...,a_m)\mapsto x_{i_1}(a_1)x_{i_2}(a_2)...x_{i_m}(a_m),}
where $x_{i_k}(a_k) = I_N+a_k E_{i_k,i_k+1}$ and $E_{i,j}$ is the matrix with 1 at the entry $(i,j)$ and 0 otherwise. Similar result also holds for $U_{>0}^-$. With this isomorphism Lusztig went on to generalize the notion of total positivity to Lie groups of arbitrary type.
In \cite{BFZ}, Berenstein et al. studied this decomposition for type $A_r$, in the context now known as \emph{cluster algebra}. They showed various relations and parametrizations using the cluster variables, in this case corresponding to determinant of the different minors. Corresponding to the canonical decomposition of $w_0$ is the parametrization using \emph{initial minors}, which are the determinants of those square sub-matrices that start from either the top row or the leftmost column. Using this parametrization, we found in \cite{FI} a family of positive principal series representations of the modular double $\cU_{q\til[q]}(\sl(N,\R))$, where the notion of the modular double was first introduced by Faddeev \cite{Fa1,Fa2} for $N=2$. These positive representations generalize the self-dual representations of $\cU_{q\til[q]}(\sl(2,\R))$ studied for example in \cite{BT,Ip,PT1}.
On the other hand, in order to study the quantum group $GL_q^+(2,\R)$ in the $C^*$-algebraic and von Neumann setting, in \cite{Ip, Pu} a quantum version of the Gauss decomposition for $GL_q(2)$ is studied, where any matrices are decomposed into product of the form
\Eq{\veca{z_{11}&z_{12}\\z_{21}&z_{22}}=\veca{u_1&0\\v_1&1}\veca{1&u_2\\0&v_2},}
where $u_iv_i=q^2v_iu_i$ are mutually commuting Weyl pair that generates the algebra of $q$-tori. In the split case, we set $|q|=1$, where
\Eq{q=e^{\pi ib^2},\tab\til[q]=e^{\pi ib^{-2}}} with $b^2\in\R\setminus\Q$, $0<b<1$. Then the Weyl pair are represented by the canonical positive essentially self-adjoint operators
\Eq{u=e^{2\pi bx},\tab v=e^{2\pi bp},} and the above decomposition gives a realization of the positive quantum group $GL_q^+(2,\R)$ where all entries and the quantum determinant are represented by positive essentially self-adjoint operators acting on $L^2(\R^2)$. Moreover, by replacing $b\to b\inv$ we obtain the representations for the modular double $GL_{q\til[q]}^+(2,\R)$. It is further shown in \cite{Ip} that $GL_q^+(2,\R)$ is the Drinfeld-Woronowicz's \emph{quantum double group} \cite{PW} over the quantum $ax+b$ group, and its harmonic analysis is studied in detail. A new Haar functional is discovered, and an $L^2$-space of ``functions" over $GL_{q\til[q]}^+(2,\R)$ is defined using this Haar functional. Then we proved that the regular representation of the modular double $U_{q\til[q]}(\sl(2,\R))$ on $L^2(GL_{q\til[q]}^+(2,\R))$ decomposes into direct integral of the positive principal series representations.
Combining the approaches above, our aim in this paper is to find the Gauss-Lusztig decomposition of the positive quantum group of higher rank, $GL_q^+(N,\R)$, in terms of the unipotent parameters $a_i$ defined above. These parameters are no longer commuting positive real numbers, and it is the goal of this paper to discover their quantum relations with each other, such that the decomposition gives precisely the definition of the quantum group $GL_q(N)$, and furthermore they are represented by positive essentially self-adjoint operators. Let us call two variables \emph{quasi-commuting} if they commute up to a power of $q^2$. In this paper we prove the following Theorem:
\begin{Thm}[Gauss-Lusztig Decomposition] The generators of the quantum group $GL_q^+(N,\R)$ can be represented by $N^2$ operators
$$\{b_{m,n}, U_k, a_{m,n}\}$$
with $1\leq n\leq m\leq N-1,1\leq k\leq N$, where each variable is positive self-adjoint operator that commutes or $q^2$-commutes with each other, so that
\begin{itemize}
\item[(1)] The variables $\{U_k, a_{m,n}\}$ generate the upper triangular quantum Borel subgroup $T_{>0}U_{>0}^+$,
\item[(2)] The variables $\{b_{m,n},U_k\}$ generate the lower triangular quantum Borel subgroup $U_{>0}^-T_{>0}$,
\item[(3)] The variables $a_{m,n}$ commute with $b_{m,n}$,
\end{itemize}
Furthermore, the Gauss-Lusztig decomposition for the other parts of the modular double $GL_{\til[q]}(N,\R)$ can be obtained by replacing all variables $\{b_{m,n}, U_k, a_{m,n}\}$ by their tilde version
\Eq{x\mapsto \til[x]:=x^{\frac{1}{b^2}}.}
\end{Thm}
As a corollary of the calculations, we also have the following results:
\begin{Thm}
\begin{itemize}
\item[(1)] There is an embedding of $GL_q^+(N,\R)$ into the algebra of $\lfloor\frac{N^2}{2}\rfloor$ $q$-tori generated by $\{u_i,v_i\}$ satisfying $u_iv_i=~q^2v_iu_i$, which are realized by
\Eq{u_i=e^{2\pi bx_i},\tab v_i=e^{2\pi bp_i}.}
\item[(2)] The quantum cluster variables $x_{ij}$, defined by the quantum determinant of the initial minors, can be represented as products of the variables $\{b_{m,n}, U_k, a_{m,n}\}$, and hence they quasi-commute with each other.
\end{itemize}
\end{Thm}
From the main Theorem, we can extend the quantum group $GL_{q\til[q]}^+(N,\R)$ into the $C^*$-algebraic setting by giving an operator norm to each element which is represented by integrals of continuous complex powers of the generators, completely analogous to the $N=2$ case. We can also give an $L^2$ completion and define the Hilbert space $L^2(GL_{q\til[q]}^+(N,\R))$. Then it is natural to conjecture its decomposition under the regular representation of the modular double $U_{q\til[q]}(\sl(N,\R))$ into the direct integral of positive principal series representations constructed in \cite{FI}, in analogy to the decompositions of $L^2(GL_{q\til[q]}^+(2,\R))$.
The Gauss decomposition for a general quantum group is definitely not new \cite{DKS, WZ}. However the usual notion in the context of $GL_q(N)$ is just decomposing the quantum group into a product of lower and upper triangular matrices, and the quantum Pl\"{u}cker relations between the coordinates are studied. Though this approach is a natural consideration, the relations involved are quite ad hoc, and furthermore it has no way to be generalized to the positive setting and its representation is rather unclear. Therefore we name our decomposition the \emph{Gauss-Lusztig decomposition} to distinguish it from the standard approach, where we decompose our quantum group into products of elementary matrices bearing a quantum variable, so that the positivity and their representations are manifest.
Finally we also remark that in \cite{BZ,FG2}, the notion of quantum cluster algebra is studied, where quasi-commuting cluster variables are considered, and the $q$-commuting relations are compatible with the algebraic framework. However its relations to the parametrization of $GL_q(N)$ is not very explicit, and its representation by the canonical $q$-tori $\{e^{2\pi bx},e^{2\pi bp}\}$ is not shown. In this paper, starting from the very definition of a quantum group, we found using new combinatorics method that these cluster variables, quasi-commuting in some complicated powers of $q^2$, are actually decomposed into simpler variables $\{b_{m,n}, U_k, a_{m,n}\}$ that commute only up to a factor of $q^2$, and explicit formula is given for the case $GL_q(N)$. The $q$-commutations we found explicitly are closely related to the Poisson structure of the cluster $\mathcal{X}$-variety considered in \cite{FG1}. We note that in this paper we only use a single choice of cluster variables given by the initial minors. A more thorough understanding of the theory of quantum cluster algebra in the context of quantum groups should be possible by also considering explicitly the quantum exchange relations to other clusters, corresponding to different parametrization of the maximal element $w_0$ explained in Theorem \ref{GLcomtrans}.
The paper is organized as follows. In Section \ref{sec:glq2} we describe in detail the Gauss decomposition for $GL_q(2)$ studied in \cite{Ip}. In Section \ref{sec:gln} we describe the Lusztig parametrization of the totally positive matrix in $GL^+(N,\R)$, and the description of the cluster variables defined in \cite{BFZ}. Then we introduce the definition of $GL_q(N)$ in Section \ref{sec:glqn}, and using certain combinatorics methods, we find in Section \ref{sec:gauss} the quantum relations between the variables of the Gauss-Lusztig decomposition. In Section \ref{sec:qtori} we constructed the representation of these quantum variables using $N^2-2$ quantum tori, and also present an example demonstrating the minimal representation using only $\lfloor \frac{N^2}{2}\rfloor$ tori. Finally using the quantum tori realization, in Section \ref{sec:posmod} we define the positive quantum group $GL_q^+(N,\R)$, and describe its relation to the modular double, and in Section \ref{sec:L2} a possible construction of an $L^2(GL_{q\til[q]}^+(N,\R))$ space.
\textbf{Acknowledgment.} This work was supported by World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan. I would like to thank my advisor Igor Frenkel for valuable comments and suggestions to this work. I would also like to thank Alexander Goncharov for pointing out Remark \ref{clusterX} in relation to the cluster $\mathcal{X}$-varieties.
\section{Gauss decomposition for $GL_q(2)$}\label{sec:glq2}
We recall the definition of $GL_q(2)$ used in \cite{FJ,Ip}. It is a rescaled version of the usual definition of $GL_q(2)$ \cite{CP}, and has an advantage of acting naturally on the standard $L^2(\R)$ space, due to the rescaled quantum determinant \eqref{z6} which resembles the classical formula without any $q$ factor.
\begin{Def}
Let $q\in\C$ which is not a root of unity. We define $GL_q(2)$ to be the bi-algebra generated by $z_{11},z_{12},z_{21}$ and $z_{22}$ subjected to the following commutation relations:
\begin{eqnarray}
\label{z1}z_{11}z_{12}&=&z_{12}z_{11},\\
z_{21}z_{22}&=&z_{22}z_{21},\\
z_{11}z_{21}&=&q^2z_{21}z_{11},\\
z_{12}z_{22}&=&q^2z_{22}z_{12},\\
z_{12}z_{21}&=&q^2z_{21}z_{12},\\
\label{z6}det_q:=z_{11}z_{22}-z_{12}z_{21}&=&z_{22}z_{11}-z_{21}z_{12}.
\end{eqnarray}
with co-product $\D$ given by
\Eq{\D(z_{ij})=\sum_{k=1,2}z_{ik}\ox z_{kj},}
and co-unit $\e$ given by
\Eq{\e(z_{ij})=\d_{ij}.}
\end{Def}
It is possible to define the antipode $\c$ by adjoining the inverse element $\det_q\inv$, giving $GL_q(2)$ a Hopf algebra structure. However we will not use the antipode in this paper.
\begin{Rem}
It is often convenient to define the matrix of generators
$$Z:=\veca{z_{11}&z_{12}\\z_{21}&z_{22}},$$
then the co-product can be rewritten as standard matrix multiplication:
\Eq{\D\veca{z_{11}&z_{12}\\z_{21}&z_{22}}=\veca{z_{11}&z_{12}\\z_{21}&z_{22}}\ox\veca{z_{11}&z_{12}\\z_{21}&z_{22}}.}
\end{Rem}
In the papers \cite{Ip,Pu}, the Gauss decomposition of $GL_q(2)$ is studied. It can be decomposed uniquely into
\Eq{\veca{z_{11}&z_{12}\\z_{21}&z_{22}}=\veca{u_1&0\\v_1&1}\veca{1&u_2\\0&v_2},}
where the Weyl pairs $\{u_i,v_i\}_{i=1,2}$ are non-commutative variables satisfying
\Eq{u_iv_i&=q^2v_iu_i,\\u_iv_j&=v_ju_i\mbox{\;\;\; for $i\neq j$.}}
Denote by $\C[\T_q]$ the algebra of quantum torus:
\Eq{\C[\T_q]:=\C\<u, v\>/(uv=q^2vu)}
Then in particular, we have an embedding of the algebra $GL_q(2)$ into the algebra of quantum tori:
\Eq{GL_q(2)\to \C[\T_q]^{\ox 2}}
Later on when we specify $|q|=1$, we will introduce a star structure so that the generators $u_i,v_i$ are self-adjoint. The representations of this algebra will be discussed in detail in Section \ref{sec:posmod}.
In order to generalize this construction to the higher rank, it turns out that it is better to write it in the form:
\Eq{\label{Gauss}\veca{1&0\\v_1&1}\veca{u_1&0\\0&1}\veca{1&0\\0&v_2}\veca{1&u_2\\0&1}}
$$=\veca{u_1&0\\v_1u_1&1}\veca{1&u_2\\0&v_2},$$
which of course still satisfies the quantum relations. Finally we note that the quantum determinant $det_q$ quasi-commutes with all other variables. It is this property that motivates us to study the Gauss decomposition for $GL_q(N)$ not using the standard coordinates, but using the ``cluster variables" which we will introduce in the next section.
\section{Parametrization of $GL^+(N,\R)$}\label{sec:gln}
In classical group theory, the total positive part $GL^+(N,\R)$ is the semi-subgroup of $GL(N,\R)$ so that all the entries are positive, and all the minors, including the determinant, are also positive. There are in general two equivalent ways to realize the total positive semi-group. In \cite{Lu}, a parametrization using the Gauss decomposition is found:
\Eq{G=U_{>0}^- T_{>0} U_{>0}^+,}
where $T_{>0}$ is the diagonal matrix with positive entries $u_i$, the positive unipotent semi-subgroup $U_{>0}^+$ (and similarly for $U_{>0}^-$) is decomposed as
\Eq{\label{u0} U_{>0}^+=\prod_{k=1}^{m}e^{a_k E_{i_k}}=\prod_{k=1}^{m}(I_N+a_k E_{i_k,i_k+1}),}
where $E_{i,i+1}$ is the matrix with 1 at the position $(i,i+1)$ and 0 otherwise, and the $i_k$'s correspond to the decomposition of the longest element $w_0$ of the Weyl group $W=S_{N-1}$:
\Eq{w_0=s_{i_1}s_{i_2}...s_{i_m}.}
Using the canonical decomposition for $w_0$:
\Eq{w_0=s_{N-1}s_{N-2}...s_2s_1s_{N-1}s_{N-2}...s_2s_{N-1}s_{N-2}...s_3...s_{N-1},}
where $s_k=(k, k+1)$ are the standard transpositions, $U_{>0}^+$ can be expressed in the form:
\footnotesize
\Eq{\label{U+}\veca{1&a_{1,1}&0&0&0\\0&1&a_{2,1}&0&0\\0&0&1&\ddots&0\\0&0&0&\ddots&a_{N-1,1}\\0&0&0&0&1}\veca{1&0&0&0&0\\0&1&a_{2,2}&0&0\\0&0&1&\ddots&0\\0&0&0&\ddots&a_{N-1,2}\\0&0&0&0&1}\cdots\veca{1&0&0&0&0\\0&1&0&0&0\\0&0&1&\ddots&0\\0&0&0&\ddots&a_{N-1,N-1}\\0&0&0&0&1}.}
\normalsize
The labeling is clear: $a_{m,n}$ is the entry at the $m$-th row and appears the $n$-th time from the left.
Similarly, $U_{>0}^-$ is given by the transpose of the form of $U_{>0}^+$, i.e.
\footnotesize
\Eq{\label{U-}\veca{1&0&0&0&0\\0&1&0&0&0\\0&0&1&0&0\\0&0&\ddots&\ddots&0\\0&0&0&b_{N-1,1}&1}\cdots\veca{1&0&0&0&0\\0&1&0&0&0\\0&b_{2,1}&1&0&0\\0&0&\ddots&\ddots&0\\0&0&0&b_{N-1,N-2}&1}\veca{1&0&0&0&0\\b_{1,1}&1&0&0&0\\0&b_{2,2}&1&0&0\\0&0&\ddots&\ddots&0\\0&0&0&b_{N-1,N-1}&1}.}
\normalsize
Under this parametrization, Berenstein et al. \cite{BFZ} studied the parametrization by cluster variables, in this case corresponds to the ``initial minors" of the matrix. These are the determinants of the square minors which start from either the top row or the leftmost column. More precisely, a matrix $g\in GL(N,\R)$ is totally positive if and only if all its initial minors are strictly positive. Furthermore, the initial minor can be expressed uniquely as a product of the parameters $a_{ij},b_{ij}$ and $u_i$, hence giving a 1-1 correspondence between the parametrizations.
In the study of the quantum Gauss decomposition, it turns out that it is just enough to look at $T_{>0}U_{>0}^+$. Let us first consider $U_{>0}^+$. Denote by $x_{ij}$, $1\leq i<j\leq N$, the determinant of the initial minor with the lower right corner at the entry $(i,j)$. Following \cite{BFZ}, there is an explicit relation between $x_{ij}$ and $a_{ij}$:
\begin{Prop} We have
\Eq{a_{i,N-j}&=\frac{x_{j,i+1}x_{j-1,i-1}}{x_{j,i}x_{j-1,i}},\\
x_{i,i+j}&=\prod_{m=1}^i \prod_{n=1}^j a_{m+n-1,n}.}
Here we denote by $x_{i,i}=x_{i,0}=x_{0,j}=1$.
\end{Prop}
The above relations can be expressed schematically by the following diagram:
$$\begin{tikzpicture}[baseline={([yshift=3ex]a33.base)}]
\node (a51) at (0, 0) {$a_{51}$};
\node (a41) at +(30: 0.866) {$a_{41}$};
\node (a31) at +(30: 1.732) {$a_{31}$};
\node (a21) at +(30: 2.598) {$a_{21}$};
\node (a11) at +(30: 3.464) {$a_{11}$};
\node (a52) at (1.5, 0) {$a_{52}$};
\node (a42) at ($(a41)+(1.5,0)$) {$a_{42}$};
\node (a32) at ($(a31)+(1.5,0)$) {$a_{32}$};
\node (a22) at ($(a21)+(1.5,0)$) {$a_{22}$};
\node (a53) at (3, 0) {$a_{53}$};
\node (a43) at ($(a42)+(1.5,0)$) {$a_{43}$};
\node (a33) at ($(a32)+(1.5,0)$) {$a_{33}$};
\node (a54) at (4.5, 0) {$a_{54}$};
\node (a44) at ($(a43)+(1.5,0)$) {$a_{44}$};
\node (a55) at (6, 0) {$a_{55}$};
\draw ([yshift=1ex]a11.north)--([xshift=1ex]a44.east)--([yshift=-1ex]a54.south)--([xshift=-1ex]a21.west)--cycle;
\node at ([yshift=3ex]a21.north) {$i$};
\node at ([yshift=3.2ex]a33.north) {$j$};
\node at (3,-0.75) {Figure 1: The cluster $x_{i,i+j}$ for $i=2,j=4$};
\end{tikzpicture}
$$
As in the $N=2$ case, we split the diagonal subgroup $T_{>0}$ into two halves:
\Eq{\label{TT}T_{>0}=T_{>0}^-T_{>0}^+:=\veca{u_1&0&0&0&0\\0&u_2&0&0&0\\0&0&\ddots&0&0\\0&0&0&u_{N-1}&0\\0&0&0&0&1}\veca{1&0&0&0&0\\0&v_1&0&0&0\\0&0&v_2&0&0\\0&0&0&\ddots&0\\0&0&0&0&v_{N-1}},}
and just consider the $v$ variables for the decomposition of the upper triangular part. Then the formulas in $T_{>0}^+U_{>0}^+$ for $a_{i,j}$ stay the same, while those for $x_{i,j}$ are modified as follows:
\Eq{x_{i,i+j}=\left(\prod_{m=1}^i \prod_{n=1}^j a_{m+n-1,n}\right)\prod_{k=1}^{i-1} v_k.}
\section{Definition of $GL_q(N)$}\label{sec:glqn}
The quantum group $GL_q(N)$ is defined by the following relations involving $GL_q(2)$:
\begin{Def} $GL_q(N)$ is the Hopf algebra generated by $\{z_{ij}\}_{i,j=1}^N$, such that for every $1\leq i<i'\leq N,1\leq j<j'\leq N$, the minor
\Eq{\veca{z_{ij}&z_{ij'}\\z_{i'j}&z_{i'j'}}}
is a copy of $GL_q(2)$, i.e. it satisfies the corresponding relations \eqref{z1}-\eqref{z6}. Furthermore, the Hopf algebra structure is given by the same classical formula:
\Eq{\D(z_{ij})&=\sum_{k=1}^N z_{ik}\ox z_{kj},\\
\e(z_{ij})&=\d_{ij}.}
Again the antipode $\c$ can be defined by adjoining $det_q\inv$ defined below, but we will not use it in the present paper.
\end{Def}
For $GL_q(N)$, the quantum determinant is again defined using the classical formula (with no $q$ involved):
\begin{Def} We define the quantum determinant as
\Eq{\label{detn} det_q = \sum_{\s\in S_N} (-1)^\s z_{1,\s(1)}...z_{N,\s(N)},}
where $S_N$ is the permutation group.
\end{Def}
Then it follows from \eqref{z6} and an induction argument that $det_q$ does not depend on the order of the row index, provided that all the monomials have the same order of row index.
As in the case of $GL_q(2)$, we can conveniently define the matrix of generators
\Eq{Z:=\big(z_{ij}\big)_{i,j=1}^N}
and simply call it the $GL_q(N)$ matrix. Then from the co-associativity of the co-product $\D$,
\Eq{\D(Z)=Z\ox Z,} we notice the following property:
\begin{Prop} If $X$ and $Y$ are $GL_q(N)$ matrices such that the generators of $X$ commute with those from $Y$, then the matrix product $G=XY$ is again a $GL_q(N)$ matrix.
\end{Prop}
Hence in order to find a Gauss decomposition $GL_q(N)=XY$ where $X$ is lower triangular and $Y$ is upper triangular, it suffices to find the corresponding matrix that satisfies the quantum relations (that any matrix can be expressed in this form is proved, for example, in \cite{DKS}). We will do this by employing the construction using the parametrizations of the totally positive matrices.
\section{Gauss-Lusztig decomposition of $GL_q(N)$} \label{sec:gauss}
Let $T^+$ and $U^+$ be given by the same matrices as in \eqref{U+} and \eqref{TT}, but instead with formal non-commuting variables $v_m, a_{mn}$ for $1\leq n\leq m\leq N-1$. We also define the cluster variables $x_{ij},1\leq i<j\leq N$ to be the quantum determinant of the initial minors of the matrix product $Z=T^+U^+$ by the determinant formula \eqref{detn}.
Then we can state our main results:
\begin{Thm}\label{main} The product $Z=T^+U^+$ is a $GL_q(N)$ matrix if and only if we have the following $q$-commutation relations between the variables given by:
\begin{itemize}
\item $a_{mn}v_m=q^2v_m a_{mn}$ for all $n$,
\item $a_{mn}a_{mn'}=q^2a_{mn'}a_{mn}$ for $n>n'$,
\item $a_{mn}a_{m-1,n'}=q^2 a_{m-1,n'}a_{mn}$ for $n\leq n'$,
\item commute otherwise.
\end{itemize}
Furthermore the variables $x_{ij}$ can be written as
\begin{eqnarray}
\label{order} x_{i,i+j}&=&\left(\prod_{m=1}^i \prod_{n=1}^j a_{m+n-1,n}\right)\prod_{k=1}^{i-1} v_k,\\
&=&(a_{11}a_{22}a_{33}...)(a_{21}a_{32}a_{43}...)...(...a_{i+j-1,j})(v_1v_2...v_{i-1})
\end{eqnarray}
in this particular order. Finally for \emph{every} $GL_q(N)$ matrix, the commutation relations between the variables $x_{ij}$ are given by
\Eq{\label{power}x_{i,i+j}x_{k,k+l}=q^{2P(i,j;k,l)} x_{l,k+l}x_{j,i+j},}
where for $j\leq l$,
\Eq{P(i,j;k,l)=\#\{m,n| l+2\leq m+n\leq k+l+1, 1\leq m\leq i, 1\leq n\leq j\}\nonumber\\
-\#\{m,n|1\leq m+n\leq i,1\leq m\leq k,1\leq n\leq l\}}
and \Eq{P(k,l;i,j)=-P(i,j;k,l).}
\end{Thm}
\begin{Cor} Let $U^-$ and $T^-$ be defined by \eqref{U-} and \eqref{TT} so that $b_{mn}$ and $u_m$ commute with $a_{mn}$ and $v_m$. Then $\{b_{mn},u_m\inv\}$ satisfies exactly the same relations as $\{a_{mn}, v_m\}$. Let $T=T^-T^+$ be the diagonal matrices with entries $T_k=u_kv_{k-1}$ for $1\leq k\leq N$, where we denote by $v_0=u_N=1$. Then the product
\Eq{GL_q(N):=Z=U^-TU^+}
gives the Gauss-Lusztig decomposition for $GL_q(N)$ parametrized by $N^2$ variables that commute up to a factor of $q^2$.
\end{Cor}
The $q$-commutation relations for $a_{mn}$ (and also $b_{mn}$) can be represented neatly by a diagram:
\Eq{\begin{tikzpicture}[baseline={([yshift=3ex]a33.base)}]
\node (a41) at (0, 0) {$a_{41}$};
\node (a31) at +(60: 1.5) {$a_{31}$};
\node (a21) at +(60: 3) {$a_{21}$};
\node (a11) at +(60: 4.5) {$a_{11}$};
\node (a42) at (1.5, 0) {$a_{42}$};
\node (a32) at ($(a31)+(1.5,0)$) {$a_{32}$};
\node (a22) at ($(a21)+(1.5,0)$) {$a_{22}$};
\node (a43) at (3, 0) {$a_{43}$};
\node (a33) at ($(a32)+(1.5,0)$) {$a_{33}$};
\node (a44) at (4.5, 0) {$a_{44}$};
\node (v1) at ($(a11)+(-1.5,0)$) {$v_1$};
\node (v2) at ($(a21)+(-1.5,0)$) {$v_2$};
\node (v3) at ($(a31)+(-1.5,0)$) {$v_3$};
\node (v4) at ($(a41)+(-1.5,0)$) {$v_4$};
\node at ($(a41)+(0,-0.5)$) {$\vdots$};
\node at ($(a42)+(0,-0.5)$) {$\vdots$};
\node at ($(a43)+(0,-0.5)$) {$\vdots$};
\node at ($(a44)+(0,-0.5)$) {$\vdots$};
\draw [-angle 90] (a21) -- (a11);
\draw [-angle 90] (a43) -- (a33);
\foreach \from/\to in {a21/a11,a31/a21,a31/a22,a32/a22,a41/a31,a41/a32,a41/a33,a42/a32,a42/a33,a43/a33}{
\draw [-angle 90] (\from) -- (\to);
}
\foreach \from/\to in {a22/a21,a32/a31,a33/a32,a42/a41,a43/a42,a44/a43}{
\draw [-angle 90]([yshift=0.8ex]\from.west) -- ([yshift=0.8ex]\to.east);
\draw [-angle 90]([yshift=-0.6ex]\from.west) -- ([yshift=-0.6ex]\to.east);
}
\foreach \from/\to in {a11/v1,a21/v2,a31/v3,a41/v4}{
\draw [-angle 90](\from) -- (\to);
}
\end{tikzpicture}}
where $u\to v$ means $uv=q^2vu$, and double arrows means it $q^2$-commutes with everything in that direction. In other words, the arrows consist of all the possible left directions, and all the north-east directions going up one level. Furthermore, note that the commutation relations for $a_{mn},v_m, u_m$ and $b_{mn'}$ are just copies of the Gauss decomposition \eqref{Gauss} for $GL_q(2)$.
\begin{Rem}\label{clusterX} It was pointed out by A. Goncharov that if we make a change of variables by taking ratios of the generators:
\Eq{a'_{m,n}=\case{a_{m,1}&n=1,\\qa_{m,n}a_{m,n-1}\inv&n>1,}}
(the $q$ factor is used to preserve positivity, cf. Section \ref{sec:posmod}) then the commutation relations among the $a'_{m,n}$ variables take a more symmetric form, represented by the diagram
\Eq{\begin{tikzpicture}[baseline={([yshift=3ex]a33.base)}]
\node (a41) at (0, 0) {$a_{41}'$};
\node (a31) at +(60: 1.5) {$a_{31}'$};
\node (a21) at +(60: 3) {$a_{21}'$};
\node (a11) at +(60: 4.5) {$a_{11}'$};
\node (a42) at (1.5, 0) {$a_{42}'$};
\node (a32) at ($(a31)+(1.5,0)$) {$a_{32}'$};
\node (a22) at ($(a21)+(1.5,0)$) {$a_{22}'$};
\node (a43) at (3, 0) {$a_{43}'$};
\node (a33) at ($(a32)+(1.5,0)$) {$a_{33}'$};
\node (a44) at (4.5, 0) {$a_{44}'$};
\node (v1) at ($(a11)+(-1.5,0)$) {$v_1$};
\node (v2) at ($(a21)+(-1.5,0)$) {$v_2$};
\node (v3) at ($(a31)+(-1.5,0)$) {$v_3$};
\node (v4) at ($(a41)+(-1.5,0)$) {$v_4$};
\node at ($(a41)+(0,-0.5)$) {$\vdots$};
\node at ($(a42)+(0,-0.5)$) {$\vdots$};
\node at ($(a43)+(0,-0.5)$) {$\vdots$};
\node at ($(a44)+(0,-0.5)$) {$\vdots$};
\node at ($(v4)+(0,-0.5)$) {$\vdots$};
\draw [-angle 90] (a21) -- (a11);
\draw [-angle 90] (a43) -- (a33);
\foreach \from/\to in {a21/a11,a11/a22,a22/a21,a31/a21,a21/a32,a32/a31,a32/a22,a22/a33,a33/a32,a41/a31,a31/a42,a42/a41,a42/a32,a32/a43,a43/a42,a43/a33,a33/a44,a44/a43}
\draw [-angle 90] (\from) -- (\to);
\draw [-angle 90](a11)--(v1);
\foreach \from/\to in {a21/v2,a31/v3,a41/v4}{
\draw [-angle 90]([yshift=0.8ex]\from.west) -- ([yshift=0.8ex]\to.east);
\draw [-angle 90]([yshift=-0.6ex]\from.west) -- ([yshift=-0.6ex]\to.east);
}
\end{tikzpicture}}
This choice of generators is closely related to the Poisson structure of the cluster $\mathcal{X}$-varieties, studied for example in \cite{FG1}.
\end{Rem}
We will use several lemmas to prove the theorem.
\begin{Lem}\label{lem-order} Assume the $q$-commutation relations in Theorem \ref{main} for $a_{mn}$ and $v_m$ hold. Then \eqref{order} holds.
\end{Lem}
\begin{proof} We use the fact that, by induction, each entry $z_{ij}$ of the upper triangular matrix has a closed form expression given by
\begin{eqnarray}
z_{i,i+j}&=&v_{i-1}\sum_{1\leq t_1<t_2<...<t_{j}\leq i+j-1} (a_{i,t_1}a_{i+1,t_2}...a_{i+j-1,t_{j}})\nonumber\\
&:=&\sum_t S_{i,t}.\label{eachterm}
\end{eqnarray}
We also have $z_{i,i}=1$ and $z_{i,i-j}=0$.
Hence the quantum initial minor $x_{i,j}$ is given by sums of products of the form \Eq{\bS_{j,t}=S_{1,t_1}S_{2,t_2}...S_{j,t_j}.}
Now using the $q$-commutation relations, which say that $a_{mn}$ commutes with $a_{m'n'}$ when both $m>m'$ and $n>n'$, we can arrange the order on each monomial $\bS_{j,t}$ so that it has a ``maximal" ordering: If the product $a_{mn}a_{m'n'}$ appears in the ordering, then either $m'=m+1$ and $n'>n$, or $m'<m$. Furthermore, if the last term in $S_{k,t}$ is $a_{m,n}$, then the term $a_{m+1,n'}$ for $n'>n$ will not appear in $S_{k+1,t}$, so that nothing can commute to the front, while we can push all the $v_m$ to the back since $v_m$ commutes with $a_{m'n}$ for $m<m'$.
This ordering is unique in the sense that for every monomial where the order in which $a_{p,*}$ appears for each fixed $p$ is the same, the corresponding maximal ordering is the same. Hence the classical calculation works and all the terms will cancel, except the one with minimal lexicographical ordering. This term is precisely
$$S_{1,t_{min}}S_{2,t_{min}}...S_{i,t_{min}},$$
where
$$S_{k,t_{min}}=v_{k-1}a_{k,1}a_{k+1,2}...a_{k+j-1,j}.$$
Again each $v_{k-1}$ in each $S_k$ commutes with all the $a$'s, so we can move them towards the back, and hence giving the expression \eqref{order}.
\end{proof}
\begin{Lem}\label{lem-power} Assume the $q$-commutation relations for $a_{mn}$ and $v_m$ hold. Then \eqref{power} holds.
\end{Lem}
\begin{proof}
Using the expression given by Lem \ref{lem-order}, we can study how $x_{ij}$ and $x_{kl}$ commute. We do this by counting how many $q$-commutations it takes for a fixed $a_{m,n}$ appearing in $x_{i,i+j}$ to travel through each variable $a_{m',n'}$ in $x_{k,k+l}$.
First, notice that $a_{m,n}$ appears in $x_{k,k+l}$ only if $1\leq n\leq l$ and $0\leq m-n\leq k-1$. Now fix $m,n$ and consider $a_{mn}$. It $q^2$-commutes with $a_{m'n'}$ in $x_{k,k+l}$ when:
\begin{itemize}
\item $q^2$: $a_{m,n'}$ with $n'<n$, hence also $1\leq n'\leq l$ and $0\leq m-n'\leq k-1$. We can rewrite this as
$$A_1=\#\{n'|\max(m+1-k,1)\leq n'\leq \min(l,n-1,m)\},$$
\item $q^{-2}$: $a_{m,n'}$ with $n'>n$, hence also $1\leq n'\leq l$ and $0\leq m-n'\leq k-1$ which reduces to
$$A_2=\#\{n'|\max(m+1-k,n+1)\leq n'\leq \min(l,m)\},$$
\item $q^2$: $a_{m-1,n'}$ with $n'\geq n$, hence
$$A_3=\#\{n'|\max(m-k,n)\leq n'\leq \min(l,m-1)\},$$
\item $q^{-2}$: $a_{m+1,n'}$ with $n'\leq n$, hence
$$A_4=\#\{\n'|\max(m-l+2,1)\leq n'\leq \min(l,n,m+1)\}.$$
\end{itemize}
Hence the amount of $q^2$ powers picked up is just the signed sum of the count above. By a case by case study, these expressions can be simplified:
$$A_3-A_2=\left\{\begin{array}{cc}1& m+n\geq l+2,n+m\leq k+l+1, n\leq l\\0& \mbox{otherwise,}\end{array}\right.$$
$$A_1-A_4=\left\{\begin{array}{cc}1&n\geq l+1, k+1\leq m+n\leq k+l\\ -1& m+n\leq k,1\leq n\leq l\\0&\mbox{otherwise.}\end{array}\right.$$
Hence, the total amount of power picked up after summing all $m,n$ is given by
\begin{eqnarray*}
&&\#\{l+2\leq m+n\leq k+l+1,n\leq l\}+\#\{k+1\leq m+n\leq l+k, l+1\leq n \}\\
&&\tab-\#\{m+n\leq k,1\leq n\leq l\},
\end{eqnarray*}
subject to $1\leq m\leq i, 1\leq n\leq j$.
Let us assume $j\leq l$. Then $n\leq j\leq l$, and the expression can be simplified to
$$\#\{l+2\leq m+n\leq k+l+1\}-\#\{m+n\leq k\}.$$
subject to $1\leq m\leq i, 1\leq n\leq j$.
This takes care of $a_{mn}$.
We still need to calculate those for $v_m$. Since there is only one $v_m$ appearing in $x_{k,k+l}$ for each $1\leq m\leq k$, we just need to count how many $a_{mn}$'s with index $m\leq k+1$ are there. Hence using the renamed $a_{m+n-1,n}$ the condition is
$$\#\{m,n|m+n\leq k, 1\leq m\leq i, 1\leq n\leq j\},$$
and this is the amount of $q^{2}$ picked up, hence canceled with the last term in the previous calculation.
Similarly considering the other direction, the amount of $q^{-2}$ picked up is
$$\#\{m,n|m+n\leq i, 1\leq m\leq k, 1\leq n\leq l\}.$$
Hence we arrive at our formula.
\end{proof}
\begin{Lem}\label{lem-power2} We have
\Eq{P(i,j;k,l)=P(i,j-1,k,l-1).}
\end{Lem}
\begin{proof} This is done by simple counting. Assume $j\leq l$. Let us compare the difference between the corresponding terms of the $P$ function. We have for the second term:
\begin{eqnarray*}
&&\#\{m,n| 1\leq m+n\leq i,1\leq m\leq k,1\leq n\leq l\} \\
&&\tab- \#\{m,n| m+n\leq i,1\leq m\leq k,1\leq n\leq l-1\}\\
&=&\#\{m,n| 1\leq m+l\leq i,1\leq m\leq k\},\\
\end{eqnarray*}
while for the first term we have
\begin{eqnarray*}
&&\#\{m,n| l+2\leq m+n\leq k+l+1, 1\leq m\leq i, 1\leq n\leq j\}\\
&&\tab-\#\{m,n| l+1\leq m+n\leq k+l, 1\leq m\leq i, 1\leq n\leq j-1\}\\
&=&\#\{m,n| l+2\leq m+n\leq k+l+1, 1\leq m\leq i, 1\leq n\leq j\}\\
&&\tab-\#\{m,n| l+2\leq m+n\leq k+l+1, 1\leq m\leq i, 2\leq n\leq j\}\\
&=&\#\{m| l+2\leq m+1\leq k+l+1, 1\leq m\leq i\}\\
&=&\#\{m| l+2\leq m+l+1\leq k+l+1, 1\leq m+l\leq i\}\\
&=&\#\{m| 1\leq m\leq k, 1\leq m+l\leq i\}.\\
\end{eqnarray*}
Hence the amounts cancel.
\end{proof}
\begin{proof}[Proof of Theorem \ref{main}] We will prove the theorem by induction. When $N=2$ it is just
$$\veca{1&0\\0&v_1}\veca{1&a_{11}\\0&1}$$
with $a_{11}v_1=q^2 v_1 a_{11}$. Hence this case holds trivially.
Assume everything hold for $\dim=N-1$.
For $\dim=N$, first we notice that $a_{N-1,N-1}$ commutes with $a_{ii}$ for $i<N-1$ by looking at the entry $z_{1,i+1}=a_{11}a_{22}...a_{ii}$, which commutes with each other by the $GL_q(2)$ relations.
Next we notice that the cluster variables for a general $GL_q(N)$ matrix depend only on the variables appearing in $T^+U^+$, since we assumed that the lower triangular matrix $U^-T^-$ commutes with $T^+U^+$. Hence the relations between $x_{i,i+j}$ which hold for $T^+U^+$ will also hold for $GL_q(N)$.
Now for a general cluster variable $x_{k,N}$ in the new rank, we know from Lemma \ref{lem-order} that $a_{N-1,N-k}$ is the only new term appearing. Hence the commutation relations between $a_{N-1,N-k}$ and $a_{i+j-1,j}$ is equivalent to the commutation relations between $x_{k,N}$ and $x_{i,i+j}$ by induction on new terms. Now consider the $(N-1)\times (N-1)$ minor corresponding to $x_{N-1,N}$. This by definition satisfies the $GL_q(N-1)$ relations, and in particular the commutation relation between
$x_{k,N}$ and $x_{i,i+j}$ should be the same as the relation between $x_{k,N-1}$ and $x_{i,i+j-1}$. However, this is precisely the statement proved in Lemma \ref{lem-power2}.
\end{proof}
The above relations can be generalized to arbitrary reduced expression for $w_0$ as follows.
Let $(a,b,c)$ and $(a',b',c')$ be positive $q$-commuting variables such that
\Eq{\label{diagram}\begin{tikzpicture}[baseline={([yshift=3ex]c.base)}]
\node (a) at (0, 0) {$a$};
\node (b) at +(50: 1.17) {$b$};
\node (c) at +(0: 1.5) {$c$};
\foreach \from/\to in {c/a, a/b}
\draw [-angle 90] (\from) -- (\to);
\end{tikzpicture}
\tab\mbox{and}\tab
\begin{tikzpicture}[baseline={([yshift=3ex]b.base)}]
\node (a) at (0, 0) {$a'$};
\node (b) at +(-50: 1.17) {$b'$};
\node (c) at +(0: 1.5) {$c'$};
\foreach \from/\to in {b/c, c/a}
\draw [-angle 90] (\from) -- (\to);
\end{tikzpicture}}
where again $u\to v$ means $uv=q^2vu$.
Then as in \eqref{u0}, the products $$x_2(a)x_1(b)x_2(c)=x_1(a')x_2(b')x_1(c')$$ form a copy of $U^+$ of the Gauss Decomposition of $GL_q(3)$ corresponding to the reduced expressions
$$w_0=s_2s_1s_2=s_1s_2s_1,$$ where
\begin{eqnarray*}
a'&=&(a+c)\inv cb=bc(a+c)\inv,\\
b'&=&a+c,\\
c'&=&(a+c)\inv ab=ba(a+c)\inv,
\end{eqnarray*}
and this map
\Eq{\phi:(a,b,c)\mapsto(a',b',c')\label{inphi}}
is an involution between $(a,b,c)\corr(a',b',c')$. In particular, we see that by applying this transformation to any three consecutive variables $a_{mn}$ corresponding to the sub-word of the form $s_is_js_i$ with $i$ adjacent to $j$, all the arrows in the diagram \eqref{diagram} are preserved. Applying this transformation, we can deduce all quantum Lusztig's variables for arbitrary reduced expression for $w_0$. Hence we can restate the commutation relations in Theorem \ref{main} as follows:
\begin{Thm}\label{GLcomtrans} Let $a_{i_n,m}$ be the coordinates of $U^+$ corresponding to the reduced expression of $w_0=s_{i_1}...s_{i_n}$. Then the product $T^+U^+$ is a $GL_q(N)$ matrix if and only if for any $|i-j|=1$, the coordinates $\{v_i,v_j,a_{i,m},a_{j,n},a_{i,k}\}$ form a copy of $GL_q(3)$, where $\{a_{i,m},a_{j,n},a_{i,k}\}$ appear in this exact order in the parametrization of $U^+$. In other words, we have
\Eq{\begin{tikzpicture}[baseline={([yshift=3ex]c.base)}]
\node (a) at (0, 0) {$a_{i,m}$};
\node (b) at +(50: 1.5) {$a_{j,n}$};
\node (c) at +(0: 1.93) {$a_{i,k}$};
\node (vi) at (-2,1.15) {$v_i$};
\node (vj) at (-2,0) {$v_j$};
\node (dot) at (-0.5,1.15) {$\cdots$};
\node (dot2) at (-0.5,0) {$\cdots$};
\foreach \from/\to in {c/a, a/b}
\draw [-angle 90] (\from) -- (\to);
\draw [-angle 90]([yshift=0.8ex]dot2.west) -- ([yshift=0.8ex]vj.east);
\draw [-angle 90]([yshift=-0.6ex]dot2.west) -- ([yshift=-0.6ex]vj.east);
\draw [-angle 90](dot) --(vi);
\end{tikzpicture}}
\end{Thm}
\section{Embedding into the algebra of quantum tori}\label{sec:qtori}
We would like to find an embedding of $GL_q(N)$ into copies of the algebra of quantum tori $\C[\T_q]$. Hence the remaining task is to find an appropriate realization of the generators $a_{mn},v_m$ using several copies of the Weyl pair $\{u,v\}$ satisfying $uv=q^2vu$.
\begin{Thm}\label{qtori1} There is an embedding of algebra $$T^+U^+\to \C[\T_q]^{\ox \frac{N^2+N-4}{2}}$$
given by
\Eq{v_m&\mapsto v_m\\
a_{mn}&\mapsto u_m \left(\prod_{k=n}^{m-1}v_{m-1,k}\right)\left(\prod_{l=1}^{n-1}v_{m,l}\right)u_{m,n}.}
where $T^+U^+$ is now realized as the algebra generated by $\{a_{mn},v_m\}$ satisfying the relations from Theorem \ref{main}, and $\C[\T_q]^{\ox \frac{N^2+N-4}{2}}$ is generated by the Weyl pairs $\{u_m,v_m\}$ and $\{u_{mn},v_{mn}\}$ for $1\leq n\leq m\leq N-1$, where we have omitted the last set of generators $\{u_{N-1,N-1},v_{N-1,N-1}\}$. (We define $u_{N-1,N-1}:=1$ in the formula).
Similarly, for $U^-T^-$ generated by $\{b_{mn},u_m\}$, we have the embedding
$$U^-T^-\to \C[\T_q]^{\ox \frac{N^2+N-4}{2}}$$
given by
\Eq{u_m&\mapsto u'_m\\
b_{mn}&\mapsto v'_m \left(\prod_{k=n}^{m-1}v'_{m-1,k}\right)\left(\prod_{l=1}^{n-1}v'_{m,l}\right)u'_{m,n}.}
where the generators $\{u',v'\}$ (with same indexing above) commute with $\{u,v\}$ used above.
Together, this gives an embedding of algebra
\Eq{GL_q(N)=U^-T^-T^+U^+\longrightarrow \C[\T_q]^{\ox N^2+N-4}.}
\end{Thm}
\begin{proof} The proof is straightforward to check, since for the $u$ variables only $u_m$ and $u_{mn}$ appear in $a_{mn}$. Hence we just need to count, at most once, how many $v_{mn}$ appears in another variables.
\end{proof}
\begin{Rem} This embedding resembles the Drinfeld double construction, which reads
\Eq{\cD(\cU_q(\fb))=\cU_q(\g)\ox \cU_q(\fh).}
With our assignment for $U^-T^-T^+U^+$, we can actually combine the diagonal variables (hence ``modding" out $\fh$) as follows:
\Eq{V_i=u'_iv_{i-1}\tab 1\leq i\leq N,}
\Eq{U_{i}=u_{i-1}={v'_i}\inv,}
(where $v_0=u'_N=1$) which gives an embedding of $GL_q(N)$ using only $(N^2-2)$ copies of $q$-tori (with inverses adjoined).
\end{Rem}
This is just one example of realizing the quantum variables where we can actually write down explicit expressions. In fact the minimal amount needed can be substantially smaller:
\begin{Thm}\label{qtori2}
The minimal amount of $q$-tori needed to realize $T^+U^+$ is given by
$\lfloor\frac{N^2}{4}\rfloor$
and the full group $GL_q(N)$ can be embedded into $\C[\T_q]^{\ox \lfloor\frac{N^2}{2}\rfloor}$.
\end{Thm}
\begin{proof}
Consider the symplectic form on the variables $\{a_{mn},u_m\}$ defined by
\Eq{\<x,y\>=\case{1&xy=q^2yx\\0&xy=yx\\-1 &xy=q^{-2}yx.}}
Then the minimal amount of $q$-tori needed to realize such relations can be found by finding the signature of this form. The skew-symmetric matrix of size $\frac{N^2+N-2}{2}$ encoding this form is actually quite simple. If we index the variables by $$a_{11},v_{1},a_{21},a_{22},v_{2},a_{31},...,$$
the upper triangular part of the matrix is given by $m+1$ consecutive 1's to the right starting at the first off diagonal entry corresponding to $a_{m*}$, truncated at the boundary, and zero otherwise. The full matrix is then obtained by anti-symmetrizing it, and we can find its kernel by elementary operations.
\end{proof}
In principle, it is possible to find the decomposition into the $q$-tori by diagonalizing the skew-symmetric matrix, corresponding to the symplectic form, into blocks $\veca{0&1\\-1&0}$ and read out the transformation.
\begin{Ex}As an example, we illustrate the cases up to $N=6$, giving the embedding of $T^+U^+$ into
$\lfloor\frac{N^2}{4}\rfloor = 9$ copies of the algebra of $q$-tori generated by $\{U_i,V_i\}$ without any powers or inverses (for $N<6$ we ignore the extra tori):
\begin{eqnarray*}
a_{11}&=&U_1\\
u_1&=&V_1\\\\
a_{21}&=&V_1U_2\\
a_{22}&=&qV_2U_2U_3\\
u_2&=&V_2\\\\
a_{31}&=&V_2U_3U_4\\
a_{32}&=&V_3U_4U_5\\
a_{33}&=&qV_4U_4U_5\\
u_3&=&V_4\\\\
a_{41}&=&V_4U_5U_6V_7V_8\\
a_{42}&=&qV_5U_5U_6V_7\\
a_{43}&=&U_3V_5U_6\\
a_{44}&=&qV_6U_6\\
u_4&=&V_6\\\\
a_{51}&=&V_6U_9\\
a_{52}&=&qV_6U_8V_9U_9\\
a_{53}&=&V_6U_7V_7V_9U_9\\
a_{54}&=&U_5V_6V_7U_8V_8V_9U_9\\
a_{55}&=&qV_8V_9U_9\\
u_5&=&V_9.\\
\end{eqnarray*}
Again the extra $q$ factors are introduced for positivity, as explained in the next section.
\end{Ex}
\begin{Con} It is possible to decompose each $a_{mn}$ into a product of \emph{single} $U_i$ and $V_i$ with the minimal amount of copies. This means that we have an embedding of $GL_q(N)$ into the \emph{polynomial} algebra generated by the minimal amount of $U_i$ and $V_i$, where the matrix entries $z_{ij}$ are expressed only in terms of polynomials of $U_i$ and $V_i$ with coefficients of the form $+q^n$.
\end{Con}
\section{Positivity and the modular double}\label{sec:posmod}
Let $q=e^{\pi ib^2}$ with $0<b^2<1, b\in\R\setminus\Q$ so that $|q|=1$ is not a root of unity. In this section we introduce the notion of the positive quantum semi-group $GL_q^+(N,\R)$ and Faddeev's notion of the modular double of $GL_q^+(N,\R)$. The basic idea is to represent the generators $z_{ij}$ in terms of positive essentially self adjoint operators acting on certain Hilbert space, where these operators are necessarily unbounded.
First we introduce the definition of an integrable representation of the canonical commutation relation defined by \cite{Sch}:
\begin{Def}\label{Sch}Let $X,Y$ be positive essentially self-adjoint operators acting on a Hilbert space $\cH$. The $q$-commutation relation $``uv=q^2vu"$ is defined to be
\Eq{u^{is}v^{it}=q^{-2st}v^{it}u^{is}}
for any $s,t\in\R$ as relations of bounded operators, where $u^{is}$ and $v^{it}$ are unitary operators on $\cH$ by the use of functional calculus.
\end{Def}
A canonical irreducible integrable representation of $uv=q^2vu$ is given by
\Eq{\label{uv}u=e^{2\pi bx}, \tab v=e^{2\pi bp},}
where $p=\frac{1}{2\pi}\del[,x]$ acting as unbounded operators on $L^2(\R)$.
On the other hand, the idea of the modular double of the Weyl pair $\{u,v\}$ is introduced by Faddeev in \cite{Fa1}. There it is suggested that for positive self-adjoint operators $u,v$ as in \eqref{uv}, one has to consider also the operators given by
\Eq{\til[u]:=u^{\frac{1}{b^2}},\tab \til[v]:=v^{\frac{1}{b^2}}}
so that \Eq{\til[u]\til[v]=\til[q]^2\til[v]\til[u],\tab \til[q]=e^{\pi ib^{-2}},}
and $\{u,v\}$ commute with $\{\til[u],\til[v]\}$ in the weak sense (the spectrum do not commute), so that the operators generated by $\{u,v,\til[u],\til[v]\}$ acting on $L^2(\R)$ is algebraically irreducible. This idea is subsequently extended to the modular double of quantum groups \cite{Fa2}.
The key technical tool in the above contexts is given by the following Lemma introduced by Volkov \cite{Vo}, and the self-adjointness is analyzed in \cite{Ru}, see also \cite{BT,Ip}:
\begin{Lem}\label{b2lem} If $u$ and $v$ are positive essentially self-adjoint operators such that $uv=q^2vu$ for $q=e^{\pi ib^2}$ in the sense above, then $u+v$ is also positive essentially self-adjoint, and we have
\Eq{(u+v)^{\frac{1}{b^2}}=u^{\frac{1}{b^2}}+v^{\frac{1}{b^2}}.}
Hence by induction, if we have $u_iu_j=q^2u_ju_i$ for every $i<j$, then the sum
$$z=\sum_i u_i$$
is positive essentially self-adjoint, and we have
\Eq{z^{\frac{1}{b^2}}=\sum_i u_i^{\frac{1}{b^2}}.}
\end{Lem}
Hence using the Gauss-Lusztig decomposition of $GL_q(N)$ defined in Theorem \ref{main}, we can define
\begin{Def} $GL_q^+(N,\R)$ is the algebra generated by positive self-adjoint operators $\{a_{mn},u_m,v_m,b_{mn}\}$ so that the $q$-commutation relations are satisfied in the sense of Definition \ref{Sch} above. These operators are acting by the Weyl pair $u=e^{2\pi bx}, v=e^{2\pi bp}$ on the Hilbert space $\cH=L^2(\R^{\lfloor\frac{N^2}{2}\rfloor})$ using Theorem \ref{qtori2}.
\end{Def}
Using Lemma \ref{b2lem}, the notion of $GL_q^+(N,\R)$ as the $q$-analogue of the classical totally positive semi-group is justified:
\begin{Cor}\label{cor1} Under the Gauss-Lusztig decomposition, the generators $z_{ij}$, as well as the initial minors $x_{ij}$ and the quantum determinant $det_q$ are represented by positive essentially self-adjoint operators.
\end{Cor}
\begin{proof} The only issue concerns the essential self-adjointness of $z_{ij}$. From the expression \eqref{eachterm}, we note that $$S_{i,t}S_{i,t'}=q^2S_{i,t'}S_{i,t},$$
whenever $S_{i,t'}$ appears later than $S_{i,t}$ in the sum. Hence using Lemma \ref{b2lem}, we conclude that each $z_{i,i+j}$ is positive and essentially self-adjoint.
\end{proof}
Let $\til[q]=e^{\pi ib^{-2}}$ and define $GL_{\til[q]}^+(N,\R)$ to be the algebra generated by the positive self-adjoint operators $\{\til[a_{mn}],\til[u_m],\til[v_m],\til[b_{mn}]\}$ where $\til[X]:=X^{\frac{1}{b^2}}$. Then the last statement of Lemma \ref{b2lem} establishes the transcendental relations between the two parts of the modular double $GL_{q\til[q]}^+(N,\R):=GL_q^+(N,\R)\ox GL_{\til[q]}^+(N,\R)$:
\begin{Thm} The generators $\til[z]_{ij}$ represented by the Gauss-Lusztig decomposition is related to $z_{ij}$ by
\Eq{\label{zz}\til[z_{ij}]=z_{ij}^{\frac{1}{b^2}},}
which is well-defined as positive essentially self-adjoint operators. Furthermore the co-product is preserved:
\Eq{(\D z_{ij})^{\frac{1}{b^2}}=\D\til[z_{ij}],}
and $z_{ij}$ commutes (weakly) with $\til[z_{ij}]$.
\end{Thm}
\begin{proof} Equation \eqref{zz} follows immediately from the proof of Corollary \ref{cor1}. The co-product is preserved because
\Eq{\D(z_{ij})=\sum_k z_{ik}\ox z_{kj},}
and $$(z_{ik}\ox z_{kj})(z_{ik'}\ox z_{k'j})=q^2(z_{ik'}\ox z_{k'j})(z_{ik}\ox z_{kj})$$
whenever $k<k'$, hence we can apply Lemma \ref{b2lem} and induction to obtain
$$\D(z_{ij})^{\frac{1}{b^2}}=\sum_k z_{ik}^{\frac{1}{b^2}}\ox z_{kj}^{\frac{1}{b^2}}=\D(\til[z_{ij}]).$$
Since $\{u,v\}$ commute weakly with $\{\til[u],\til[v]\}$, the last statement follows by the Gauss-Lusztig decomposition.
\end{proof}
\section{$GL_{q\til[q]}^+(N,\R)$ in the $L^2$ setting}\label{sec:L2}
In the final section we generalize the approach from \cite{Ip} to higher rank, where the $L^2$-space for $GL_{q\til[q]}^+(2,\R)$ is introduced. The $L^2$-norm comes from the classical counterpart of the Haar measure. In the classical case, under the decomposition given by \eqref{Gauss}, the Haar measure induced on the coordinates is given by
\Eq{dg=\frac{du_1}{u_1}dv_1\frac{dv_2}{v_2}du_2.}
When the original positive variable has $dx$ as the measure, the Mellin transform gives
\Eq{\|f(x)\|_{dx}^2=\left\|\int_{\R+i0} f(s)x^{is}ds\right\|^2=\int_\R |f(s+\frac{i}{2})|^2 ds.\label{L2norm}}
Similarly, when the measure of the original positive variable is given by $\frac{dy}{y}$, the Mellin transformed measure remains unchanged:
\Eq{\|f(y)\|_{\frac{dy}{y}}^2=\left\|\int_\R f(s)y^{is}ds\right\|^2 = \int_\R |f(s)|^2ds.\label{L2norm2}}
It is shown in \cite{Ip} that in the quantum setting, using certain GNS representation of $GL_{q\til[q]}^+(2,\R)$ in the $C^*$-algebraic setting, there exists a Haar weight which induces an $L^2$-norm on entire functions with 4 \emph{classical} coordinates $f(u_1,v_1,u_2,v_2)$, where the right hand side of \eqref{L2norm} is replaced by
\Eq{\int_\R \left|f(s+\frac{iQ}{2})\right|^2 ds}
with $Q=b+b\inv$, while the right hand side of \eqref{L2norm2} remains unchanged.
For higher rank, we notice that the Haar measure for the usual totally positive parametrization
\Eq{U_{>0}^-(b_{ij})T_{>0}(u_{i})U_{>0}^+(a_{ij})}
is given by
\Eq{\prod_{k=1}^N\frac{du_k}{u_k}\prod_{1\leq j\leq i\leq N-1}(a_{ij}b_{ij})^{N-1-i}da_{ij}db_{ij}.}
This is obtained by calculating the Jacobian of the change of variables \eqref{eachterm} with coordinates of the usual Gauss decomposition, and using the fact that the Haar measure induces a multiplicative measure on the diagonal matrices, and the standard $L^2$ measure on the upper/lower triangular unipotent matrices.
Now let us introduce a change of variables for our cluster variables:
\Eq{X_{ij}:=\left\{\begin{array}{cc}x_{ii}x_{i-1,i-1}\inv&i=j\\x_{ij}x_{ii}\inv&i< j\\x_{ij}x_{jj}\inv&i> j,\end{array}\right.}
i.e. we use only $u_i$ from the diagonal matrices, and the cluster variables of the upper/lower triangular unipotent matrices in terms of products of $a$'s or $b$'s only. Then we have
\begin{Prop} The (classical) Haar measure on $GL^+(N,\R)$ is given by:
\Eq{\left(\prod_{i,j=1}^{N-1}\frac{dX_{ij}}{X_{ij}}\right)\left(\prod_{k=1}^{N-1}dX_{N,k}dX_{k,N}\right) \frac{dX_{NN}}{X_{NN}}.}
\end{Prop}
Hence following the idea in the case of $GL_{q\til[q]}^+(2,\R)$ we can define $C_\oo(GL_{q\til[q]}^+(N,\R))$ and $L^2(GL_{q\til[q]}^+(N,\R))$ as follow:
\begin{Def} The $C^*$-algebra $C_\oo(GL_{q\til[q]}^+(N,\R))$ is the norm closure of operators of the form:
\Eq{F:=\int_{\R^{N^2}} f(s_{11},...,s_{NN})\prod_{m,n=1}^{N}(X_{mn})^{ib\inv s_{mn}} \prod_{m,n=1}^N ds_{mn},}
where $f(s_{11},...,s_{NN})$ are smooth analytic rapidly decreasing functions in each variable $s_{ij}\in~\R$, each $X_{mn}^{ib\inv s_{mn}}$ is realized as a unitary operator defined by the formula in Theorem \ref{qtori2}, and the $C^*$-norm is defined as the operator norm. Furthermore, the $C^*$-algebra corresponding to difference choices of realization of $X_{mn}$ are isomorphic.
\end{Def}
\begin{Rem} It will be interesting to put $C_\oo(GL_{q\til[q]}^+(N,\R))$ in the context of locally compact quantum group, for example in the sense of \cite{KV}. Although it is known in the case when $N=2$ \cite{Ip}, in general it is not clear how the co-product $\D$ defined on $z_{ij}$ translates to the generators $X_{mn}$, hence the main difficulty will be to show the density conditions involving $\D$.
\end{Rem}
For the sake of harmonic analysis, we can also define an $L^2$-space where it does not depend on the choice of embedding of $GL_q^+(N,\R)$ at all.
\begin{Def}
We define $L^2(GL_{q\til[q]}^+(N,\R))\simeq L^2(\R^{N^2})$ by giving an $L^2$-norm for the (rapidly decreasing entire) functions $F$:
\Eq{\|F\|^2 := \int_{\R^{N^2}}\left|f(s_{ij}+(\d_{i,N}-\d_{j,N})^2\frac{iQ}{2})\right|^2 \prod_{i,j=1}^N ds_{ij}.}
and taking the $L^2$-completion.
\end{Def}
A natural class of representations for split real quantum groups $U_{q\til[q]}(\g_\R)$ for arbitrary type simple Lie algebra $\g$, called the positive principal series representations, generalizing the $\sl(2,\R)$ case is introduced in \cite{FI} and constructed in \cite{Ip2,Ip3}. It is conjectured in \cite{PT1} and shown in \cite{Ip} that the left and right regular representations of $U_{q\til[q]}(\gl(2,\R))$ acting on $L^2(GL_{q\til[q]}^+(2,\R))$ naturally decompose as a direct integral of tensor product of the positive representations $\cP_{\l,s}$:
\Eq{L^2(GL_{q\til[q]}^+(2,\R))\simeq \int_\R^\o+ \int_{\R_+}^\o+\cP_{\l,s}\ox\cP_{\l,-s}d\mu(\l)ds}
where $\mu(\l)$ is expressed in terms of the quantum dilogarithm function. Hence analogous to the harmonic analysis of $L^2(GL_{q\til[q]}^+(2,\R))$, it is natural to ask the following question:
\begin{Con} Do the left and right regular representations of $U_{q\til[q]}(\gl(N,\R))$ on the $L^2(GL_{q\til[q]}^+(N,\R))$ space decompose as a direct integral of tensor product of the positive principal series representations of $U_{q\til[q]}(\gl(N,\R))$?
Furthermore, can analogous statements be defined for $U_{q\til[q]}(\g_\R)$ for arbitrary type simple Lie algebra $\g$?
\end{Con}
|
\section{Introduction}
Recent observations have shown the ubiquitous presence of propagating magnetohydrodynamic (MHD) Alfv\'enic waves in the solar atmosphere. For example, Alfv\'enic transverse waves propagating in magnetic waveguides of the solar corona were first observed by \citet{tomczyk07} and \citet{tomczyk09} using the Coronal Multichannel Polarimeter (CoMP), and more recently by \citet{mcintosh2011} using SDO/AIA. In chromospheric spicules, the presence of Doppler oscillations is known for more than 40 years \citep[see the review by][]{temuryreview}. Recent observations of Alfv\'enic transverse waves in spicules have been reported by, e.g, \citet{depontieu07}, \citet{temuryspicules}, \citet{kim}, \citet{he1,he2}, \citet{okamotodepontieu}. In addition, propagating transverse waves in thin threads of solar prominences have been observed \citep[e.g.,][]{lin07,lin09} and Alfv\'enic waves in bright points have been reported \citep{jess2009}. The role and implications of the observed Alfv\'enic waves for the heating of the solar atmosphere have been discussed by, e.g., \citet{robertus07,depontieu07,mcintosh2011,cargillineke}.
Based on MHD wave theory, a number of works have interpreted the observed waves as propagating kink MHD waves \citep[e.g.,][]{robertus07,tom08,lin09,pascoe1,pascoe2,TGV,VTG,verthspicule,solerspatial,resonantflow,stratified}. Kink waves are transverse waves with mixed fast MHD and Alfv\'enic properties \citep[see, e.g.,][]{edwinroberts,goossens09}. In thin magnetic tubes kink waves are highly Alfv\'enic because their dominant restoring force is magnetic tension \citep{goossens09}. It has been shown that resonant absorption, caused by plasma inhomogeneity in the direction transverse to the magnetic field, is a natural and efficient damping mechanism for kink waves \citep[see the recent reviews by][]{goossens06,goossens08,goossens11}. In magnetic waveguides resonant absorption transfers wave energy from transverse kink motions to azimuthal motions localized in the inhomogeneous part of the waveguide. This process has been studied numerically by \citet{pascoe1,pascoe2} in the case of driven kink waves in coronal waveguides.
Using analytical theory based on the thin tube and thin boundary approximations, \citet[hereafter TGV]{TGV} obtained that the damping length due to resonant absorption is inversely proportional to the wave frequency. Therefore, it was predicted that high-frequency kink waves become damped in length scales shorter than low-frequency waves. \citet{VTG} showed that this result is consistent with the CoMP observations of damped coronal waves \citep{tomczyk07,tomczyk09}. Subsequent investigations have extended the original work by TGV by incorporating effects not considered in their paper. For example, \citet{resonantflow} took the presence of flow into account, and \citet{stratified} studied the influence of longitudinal density stratification. Both works concluded that the damping length remains inversely proportional to the frequency when flows and longitudinal stratification are present.
In TGV and subsequent works cited above, the plasma is assumed fully ionized. However, the plasma in the cooler parts of the solar atmosphere is only partially ionized as, e.g., in the chromosphere or in prominences. This fact raises the relevant question on whether the damping length remains inversely proportional to the frequency when the plasma is partially ionized or, on the contrary, this dependence is modified by the effect of ion-neutral collisions. The effect of partial ionization on the damping of Alfv\'en waves has been investigated in a large number of papers in different contexts. Some examples are the works by, e.g., \citet{hearendel,depontieu98,pecseli,depontieu2001,forteza07,solerpartial,marc2010,temury,temuryhelium}. However, these works only focus on the role of partial ionization for the damping and do not consider resonant absorption, which is a basic and unavoidable phenomenon when plasma and/or magnetic inhomogeneity is present.
To our knowledge, the first attempt to study resonant waves in partially ionized plasmas was by \citet{solerpartialres}. These authors used the single-fluid approximation \citep[see, e.g.,][]{brag} to investigate standing resonant kink waves in a model of a partially ionized prominence thread. Subsequently, this first investigation was extended in \citet[hereafter SOB]{solerspatial} to the case of propagating waves. The use of the single-fluid approximation, as in \citet{solerpartialres} and SOB, seems reasonable as the expected values of the collision frequencies in the solar atmosphere are much larger than the observed wave frequencies. However, the single-fluid approximation misses important effects when small length scales and/or high frequencies are involved. In such cases, the multifluid description is a more suitable approach \citep[see, e.g.,][]{temury,temuryhelium}. For resonant waves, perturbations develop very small length scales in the vicinity of the resonance position \citep[see, e.g.,][]{tirry,rudermanwright,vasquez,terradas2006}, where multifluid effects may play a relevant role.
Here we perform a general description of propagating resonant Alfv\'enic waves in partially ionized plasmas using the multifluid treatment. Our work extends the investigation of SOB by considering arbitrary values of the collision frequencies and by performing a more in-depth analysis of the resonant process. Section~\ref{sec:basic} contains a description of the equilibrium configuration and the basic equations. In Section~\ref{sec:resonant} the behavior of wave perturbations around the Alfv\'en resonance position are investigated and jump relations for the perturbations are derived. Later, in Section~\ref{sec:tube} we perform an application to resonant kink waves in straight tubes and obtain expressions for the damping lengths due to resonant absorption and ion-neutral collisions. The approximate analytical theory is complemented with fully numerical eigenvalue computations. In Section~\ref{sec:chromos} the implications of our results for the particular case of chromospheric waves is discussed. Finally, we give our main conclusions in Section~\ref{sec:con}.
\section{Equilibrium and basic equations}
\label{sec:basic}
We consider a partially ionized plasma composed of ions, electrons, and neutrals. We use cylindrical coordinates, namely $r$, $\varphi$, and $z$ for the radial, azimuthal, and longitudinal coordinates. The medium is permeated by an equilibrium magnetic field, $\bf B$, of the form
\begin{equation}
{\bf B} = B_\varphi \hat{e}_\varphi + B_z \hat{e}_z.
\end{equation}
In general, both azimuthal, $B_\varphi$, and longitudinal, $B_z$, components are functions $r$. The equations governing the dynamics of a magnetized multifluid plasma are discussed in classical works as, e.g., \citet{cowling} and \citet{brag}, or more recently in \citet{birk} and \citet{pinto}. Recent investigations which make extensive use of the multifluid description in the context of MHD waves are \citet{temury,temuryhelium}. We refer the reader to these two papers for details about the derivation of the equations. In brief, a partially ionized multifluid plasma is governed by the equations of the different species, which contain terms that couple the various fluids by means of collisions.
Here we study linear perturbations superimposed on the equilibrium state. Due to the very small momentum of electrons, we neglect the collisions of electrons with ions and neutrals. In addition, we consider a magnetically dominated plasma and neglect the gas pressure of ions compared to the Lorentz force, and the gas pressure of neutrals compared to the collisional friction with ions. This simplification neglects the plasma displacement along the magnetic field direction and removes the slow or cusp continuum from the equilibrium. Thus, in the present work we focus on resonance absorption in the Alfv\'en continuum only. The inclusion of gas pressure is a subject for future works. Under these conditions, the basic equations of our investigation are
\begin{eqnarray}
{\rho_{\rm i}} \frac{\partial {\bf v}_{\rm i}}{\partial t} &=& \frac{1}{\mu} \left( \nabla \times {\bf b} \right) \times {\bf B} - {\rho_{\rm i}} {\rho_{\rm n}} \gamma_{\rm in} \left( {\bf v}_{\rm i} - {\bf v}_{\rm n} \right), \label{eq:momion} \\
{\rho_{\rm n}} \frac{\partial {\bf v}_{\rm n}}{\partial t} &=& - {\rho_{\rm i}} {\rho_{\rm n}} \gamma_{\rm in} \left( {\bf v}_{\rm n} - {\bf v}_{\rm i} \right), \label{eq:momneu} \\
\frac{\partial {\bf b}}{\partial t} &=& \nabla \times \left({\bf v}_{\rm i} \times {\bf B} \right), \label{eq:induct}
\end{eqnarray}
where ${\bf v}_{\rm i}$ and ${\bf v}_{\rm n}$ are the velocities of ions and neutrals, respectively, ${\bf b}$ is the magnetic field perturbation, ${\rho_{\rm i}}$ and ${\rho_{\rm n}}$ are the densities of the ion and neutral fluids, respectively, $\mu$ is the magnetic permittivity, and $\gamma_{\rm in}$ is the ion-neutral collision rate coefficient per unit mass. Instead of using $\gamma_{\rm in}$, in the remaining of this paper we use the ion-neutral collision frequency, $\nu_{\rm in}$, which has a more obvious physical meaning. The ion-neutral collision frequency is defined as
\begin{equation}
\nu_{\rm in} = \rho_{\rm i} \gamma_{\rm in}.
\end{equation}
Equations~(\ref{eq:momion}) and (\ref{eq:momneu}) are the linearized momentum equations of ions and neutrals, respectively. Equation~(\ref{eq:induct}) is the linearized induction equation, in which we have omitted all nonideal, diffusion terms as, e.g., magnetic resistivity. We do so because the goal of the present paper is to assess the particular role of ion-neutral collisions on the resonance absorption process. The effects of other dissipative mechanisms as, e.g, resistivity or viscosity, have been extensively studied in the existing literature \citep[see the recent review by][]{goossens11}
In this work the equilibrium quantities are functions of $r$ alone, so that the equilibrium is uniform in both azimuthal and longitudinal directions. Hence we can write the perturbed quantities proportional to $\exp \left( i m \varphi + i k_z z \right)$, where $m$ and $k_z$ are the azimuthal and longitudinal wavenumbers, respectively. We express the temporal dependence of the perturbations as $\exp \left( - i \omega t \right)$, with $\omega$ the frequency. Since there are no equilibrium flows, the Lagrangian displacement of ions, ${\bf \xi}_{\rm i}$, and neutrals, ${\bf \xi}_{\rm n}$, are
\begin{equation}
{\bf \xi}_{\rm i} = \frac{i}{\omega} {\bf v}_{\rm i}, \qquad {\bf \xi}_{\rm n} = \frac{i}{\omega} {\bf v}_{\rm n}.
\end{equation}
From Equation~(\ref{eq:momneu}) it is straightforward to derive the relation between ${\bf \xi}_{\rm n}$ and ${\bf \xi}_{\rm i}$, namely
\begin{equation}
{\bf \xi}_{\rm n} = \frac{i\nu_{\rm in} }{\omega + i\nu_{\rm in} } {\bf \xi}_{\rm i}. \label{eq:xiixin}
\end{equation}
Equation~(\ref{eq:xiixin}) informs us that ${\bf \xi}_{\rm n}$ and ${\bf \xi}_{\rm i}$ are related to each other. The factor of proportionally is a complex function of the wave frequency, $\omega$, and the ion-neutral collision frequency, $\nu_{\rm in}$. Hence, there is a phase difference between the motions of ions and neutrals. Physically, this phase difference can be interpreted as the ``time delay'' between the motions of ions and the corresponding reaction of neutrals. Thus, for $\nu_{\rm in} \gg \omega$, ${\bf \xi}_{\rm n} = {\bf \xi}_{\rm i}$ and the phase difference is zero, i.e., ions and neutrals behave as a single fluid. On the contrary, for $\nu_{\rm in} \ll \omega$ ions and neutrals become decoupled from each other, i.e., the collisionless case. In the remaining of this paper, we call the limits $\nu_{\rm in} \gg \omega$ and $\nu_{\rm in} \ll \omega$ the single-fluid limit and the collisionless limit, respectively.
Although we perform a general analysis for arbitrary $\nu_{\rm in}$ and $\omega$, we must note that $\nu_{\rm in} \gg \omega$ is the realistic situation according to the observed frequencies of MHD waves in the solar atmosphere and the expected values of the collision frequency \citep[see, e.g.,][]{depontieu98, depontieu2001}. Thus, the case $\nu_{\rm in} \gg \omega$ will receive special attention.
\section{General expressions for resonant Alfv\'en waves}
\label{sec:resonant}
In this section we study the properties of resonant MHD waves in the Alfv\'en continuum. We follow the notation used in, e.g., \citet{SGH91,goossens92,goossens95}. We combine Equations~(\ref{eq:momion})--(\ref{eq:induct}) to arrive at two coupled equations for the radial component of the Lagrangian displacement of ions, $\xi_{{\rm i}r} = i {\bf v_{{\rm i}}} \cdot \hat{e}_r / \omega$, and the Eulerian perturbation of the total pressure, $P={\bf B} \cdot {\bf b} / \mu$, namely
\begin{eqnarray}
\mathcal{D} \frac{{\rm d} \left( r \xi_{{\rm i}r} \right)}{{\rm d} r} &=& \mathcal{C}_1 r \xi_{{\rm i}r} - \mathcal{C}_2 r P, \label{eq:bas1} \\
\mathcal{D} \frac{{\rm d} P}{{\rm d} r} &=& \mathcal{C}_3 \xi_{{\rm i}r} - \mathcal{C}_1 P,\label{eq:bas2}
\end{eqnarray}
with
\begin{eqnarray}
\mathcal{D} &=& {\rho_{\rm i}} v_{\mathrm{Ai}}^2 \left( \tilde{\omega}^2 - \omega_{\rm A}^2 \right), \label{eq:defd}\\
\mathcal{C}_1 &=& \frac{2}{\mu} \frac{B_\varphi^2}{r} \tilde{\omega}^2 - v_{\mathrm{Ai}}^2 \frac{2}{\mu} \frac{m}{r^2} f_{\rm B} B_\varphi, \label{eq:c1} \\
\mathcal{C}_2 &=& \tilde{\omega}^2 - v_{\mathrm{Ai}}^2 \left( \frac{m^2}{r^2} + k_z^2 \right), \label{eq:c2} \\
\mathcal{C}_3 &=& \mathcal{D} \left[ {\rho_{\rm i}} v_{\mathrm{Ai}}^2 \left( \tilde{\omega}^2 - \omega_{\rm A}^2 \right) + \frac{2}{\mu} B_\varphi \frac{{\rm d}}{{\rm d} r} \left( \frac{B_\varphi}{r} \right) \right] \nonumber \\
&+& \frac{4}{\mu^2} \frac{B_\varphi^4}{r^2}\tilde{\omega}^2 - \frac{4}{\mu} {\rho_{\rm i}} v_{\mathrm{Ai}}^2 \omega_{\rm A}^2 \frac{B_\varphi^2}{r^2} \label{eq:c3},
\end{eqnarray}
where $v_{\mathrm{Ai}}^2 = \frac{B^2}{\mu {\rho_{\rm i}}}$ is the square of the Alfv\'en velocity of the ions, with $B^2 = B_\varphi^2 + B_z^2$. In addition we have defined
\begin{eqnarray}
f_{\rm B} &=& \frac{m}{r}B_\varphi + k_z B_z, \\
\omega_{\rm A}^2 &=& \frac{1}{\mu {\rho_{\rm i}}} f_{\rm B}^2, \\
\tilde{\omega}^2 &=& \omega^2 \frac{\omega + i \left( 1 + \alpha \right) \nu_{\rm in}}{\omega + i\nu_{\rm in}}, \label{eq:omegat} \\
\alpha &=& \frac{{\rho_{\rm n}}}{{\rho_{\rm i}}}.
\end{eqnarray}
In these expressions, $\omega_{\rm A}$ is the local Alfv\'en frequency for the ions, $\tilde{\omega}$ is the effective or modified frequency due to collisions, and $\alpha$ indicates the plasma ionization degree.
\subsection{Modified Alfv\'en continuum}
Equations~(\ref{eq:bas1}) and (\ref{eq:bas2}) are singular when $\mathcal{D} = 0$. As the equilibrium quantities are functions of $r$, the position of the singularity is mobile and depends on the frequency, $\omega$. The whole set of frequencies satisfying $\mathcal{D} = 0$ at some $r$ form a continuum of frequencies called the Alfv\'en continuum \citep[see, e.g.,][]{appert}. In the ideal, fully ionized case the condition $\mathcal{D} = 0$ is satisfied where the frequency, $\omega$, matches the local Alfv\'en frequency, $\omega_{\rm A}$.
We investigate how the Alfv\'en continuum is affected by ion-neutral collisions. The equation $\mathcal{D} = 0$, which $\mathcal{D}$ defined in Equation~(\ref{eq:defd}), can be rewritten as a third-order polynomial in $\omega$, namely
\begin{equation}
\omega^3 + i \left( 1 + \alpha \right) \nu_{\rm in} \omega^2 - \omega_{\rm A}^2 \omega - i \nu_{\rm in} \omega_{\rm A}^2 = 0. \label{eq:continuum}
\end{equation}
Equation~(\ref{eq:continuum}) describes the Alfv\'en continuum modified by ion-neutral collisions. For $\nu_{\rm in} = 0$ we recover the ideal Alfv\'en continuum in a fully ionized plasma, i.e., $\omega = \pm \omega_{\rm A}$. Note that Equation~(\ref{eq:continuum}) and so the continuum frequencies are independent of $m$.
To shed light on the effect of ion-neutral collisions, it is convenient to rewrite Equation~(\ref{eq:continuum}) in the following form
\begin{equation}
\omega^2 - \omega_{\rm A}^2 \frac{1 + i \frac{\nu_{\rm in}}{\omega} }{1 + i \left( 1 + \alpha \right) \frac{\nu_{\rm in}}{\omega}} = 0.
\end{equation}
This enables us to easily evaluate the continuum frequencies at the limits $\nu_{\rm in} \gg \omega$ and $\nu_{\rm in} \ll \omega$, namely
\begin{equation}
\omega \approx \left\{
\begin{array}{lll}
\pm \omega_{\rm A} \left( 1+\alpha \right)^{-1}, & \textrm{for} & \nu_{\rm in} \gg \omega, \\
\pm \omega_{\rm A}, & \textrm{for} & \nu_{\rm in} \ll \omega.
\end{array} \right.
\end{equation}
In the limit $\nu_{\rm in} \ll \omega$ we obtain again the continuum frequencies of the ideal, collisionless case. In the limit $\nu_{\rm in} \gg \omega$, the continuum frequencies are modified by the plasma ionization degree but the frequencies remain real, so that the position of the singularity of Equations~(\ref{eq:bas1}) and (\ref{eq:bas2}) is shifted with respect to the collisionless case.
The limits $\nu_{\rm in} \gg \omega$ and $\nu_{\rm in} \ll \omega$ are extreme cases. Now, we explore the continuum frequencies for arbitrary $\nu_{\rm in}$ to have a complete picture of the role of ion-neutral collisions. For arbitrary $\nu_{\rm in}$ we find an approximate solution to Equation~(\ref{eq:continuum}) by assuming that the effect of collisions on the continuum is to produce a weak damping of the continuum modes. This means that the continuum frequencies are assumed complex and that their real parts are taken the same as in the collisionless case. Thus we write $\omega \approx \pm \omega_{\rm A} + i \gamma$, with $\gamma \ll \omega_{\rm A}$, and put this expression in Equation~(\ref{eq:continuum}). After neglecting terms with $\mathcal{O} \left( \gamma^2 \right)$ we obtain the modified Alfv\'en continuum frequencies, namely
\begin{equation}
\omega \approx \pm \omega_{\rm A} - i \frac{1}{2} \alpha \nu_{\rm in}. \label{eq:continuum2}
\end{equation}
Equation~(\ref{eq:continuum2}) shows that the continuum frequencies in a partially ionized plasma are complex. Based on this result we anticipate that waves driven with a real frequency $\omega$ will not produce true singularities in Equations~(\ref{eq:bas1}) and (\ref{eq:bas2}) since the Alfv\'en continuum frequencies are in the complex plane. This means that the singularity is removed in a partially ionized plasma even in the absence of additional dissipative mechanisms as, e.g., resistivity or viscosity.
As the two solutions given in Equation~(\ref{eq:continuum2}) are complex, the third solution to Equation~(\ref{eq:continuum}) must be purely imaginary. So, we write $\omega \approx i \gamma$ and neglect terms with $\mathcal{O} \left( \gamma^2 \right)$ to obtain an expression for the third solution, namely
\begin{equation}
\omega \approx - i \nu_{\rm in},
\end{equation}
which corresponds to a purely damped, nonpropagating mode whose damping rate is given by the collision frequency. The presence of this third solution is not relevant for our subsequent analysis.
\subsection{Behavior of perturbations around the resonance position}
Here our purpose is to assess the behavior of the perturbed quantities around the position of the ideal Alfv\'en resonance. We denote by $r_{\mathrm{A}}$ the ideal Alfv\'en resonance position. The value of $r_{\mathrm{A}}$ is obtained from the ideal resonance condition, i.e., $\omega^2 = \omega_{\rm A}^2 (r_{\mathrm{A}})$, where $\omega$ is assumed to be real. Around $r=r_{\mathrm{A}}$ the behavior of the perturbations is dominated by the resonance. At $r=r_{\mathrm{A}}$ ion-neutral collisions become relevant to remove the singularity of the coefficient $\mathcal{D}$ in Equations~(\ref{eq:bas1}) and (\ref{eq:bas2}), while their effect on the rest of coefficients is minor. This means that around $r=r_{\mathrm{A}}$ we can take the ideal expressions of coefficients $\mathcal{C}_1$, $\mathcal{C}_2$, and $\mathcal{C}_3$, and only keep the terms related to collisions in the expression of coefficient $\mathcal{D}$.
We define the new variable $s=r-r_{\mathrm{A}}$. The ideal coefficients $\mathcal{C}_1$, $\mathcal{C}_2$, and $\mathcal{C}_3$ at $s=0$ become the following constant terms,
\begin{eqnarray}
\mathcal{C}_1 &\approx& - \frac{2}{\mu} \frac{B_\varphi B_z}{r_{\mathrm{A}}} \frac{f_{\rm B} g_{\rm B}}{\mu {\rho_{\rm i}}}, \label{eq:c1s0} \\
\mathcal{C}_2 &\approx& - \frac{g_{\rm B}^2}{\mu {\rho_{\rm i}}},\label{eq:c2s0} \\
\mathcal{C}_3 &\approx& - \frac{4}{\mu^2} \frac{B_\varphi^2 B_z^2}{r_{\mathrm{A}}^2} \frac{f_{\rm B}^2}{\mu {\rho_{\rm i}}},\label{eq:c3s0}
\end{eqnarray}
with $g_{\rm B} = \frac{m}{r_{\mathrm{A}}} B_z - k_z B_\varphi$. All quantities in Equations~(\ref{eq:c1s0})--(\ref{eq:c2s0}) have to be evaluated at $s=0$. Then we can rewrite Equations~(\ref{eq:bas1}) and (\ref{eq:bas2}) as
\begin{eqnarray}
\mathcal{D}\frac{{\rm d} \xi_{{\rm i}r} }{{\rm d} s} &\approx& \frac{g_{\rm B}}{\mu {\rho_{\rm i}}} \mathcal{C}_{\rm A}, \label{eq:bas1s0}\\
\mathcal{D}\frac{{\rm d} P}{{\rm d} s} &\approx& \frac{2}{\mu} \frac{B_\varphi B_z}{r_{\mathrm{A}}}\frac{f_{\rm B} }{\mu {\rho_{\rm i}}}\mathcal{C}_{\rm A}, \label{eq:bas2s0}\\
\mathcal{D}\frac{{\rm d} \mathcal{C}_{\rm A}}{{\rm d} s} &\approx& 0,
\end{eqnarray}
where $ \mathcal{C}_{\rm A}$ is defined as
\begin{equation}
\mathcal{C}_{\rm A} = g_{\rm B} P - \frac{2}{\mu} \frac{B_\varphi B_z}{r_{\mathrm{A}}} f_{\rm B} \xi_{{\rm i}r}. \label{eq:conservation}
\end{equation}
The quantity $ \mathcal{C}_{\rm A}$ is approximately conserved across the resonance position. This is the same conservation law as that in ideal and dissipative MHD \citep[e.g.,][]{SGH91,goossens95,tirry}. The Lagrangian displacement perpendicular to the magnetic field lines, $\xi_{{\rm i}\perp} = \left( \xi_{{\rm i}\varphi} B_z - \xi_{{\rm i}z} B_\varphi \right) / B$, is related to $ \mathcal{C}_{\rm A}$ as
\begin{equation}
\mathcal{D} \xi_{{\rm i}\perp} = i \frac{B}{\mu {\rho_{\rm i}}} \mathcal{C}_{\rm A}. \label{eq:xiperp}
\end{equation}
Next we approximate the coefficient $\mathcal{D}$ by its first-order Taylor polynomial around $s=0$. The linear expansion of $\mathcal{D}$ is approximately valid in the interval $-s_{\rm A} <s < s_{\rm A}$, with $s_{\rm A}$ satisfying the relation $s_{\rm A} \ll \left| (\omega_{\rm A}^2)' / (\omega_{\rm A}^2)'' \right|$, where the prime denotes radial derivative. The resulting expression is
\begin{equation}
\mathcal{D} \approx {\rho_{\rm i}} v_{\mathrm{Ai}}^2 \Delta_{\rm A} \left( \Lambda + s \right), \label{eq:ds0}
\end{equation}
with
\begin{equation}
\Delta_{\rm A} = \frac{{\rm d}}{{\rm d} s} \left( \omega^2 - \omega_{\rm A}^2 \right),
\end{equation}
and $\Lambda = \Lambda_{{\rm R}} + i \Lambda_{{\rm I}} $, where $\Lambda_{{\rm R}}$ and $\Lambda_{{\rm I}}$ are the real and imaginary parts of $\Lambda$, respectively, given by
\begin{equation}
\Lambda_{{\rm R}} = \frac{\omega^2}{\Delta_{\rm A}} \frac{\alpha \nu_{\rm in}^2}{\omega^2 + \nu_{\rm in}^2}, \qquad \Lambda_{{\rm I}} = \frac{\omega^2}{\Delta_{\rm A}} \frac{\alpha \omega \nu_{\rm in}}{\omega^2 + \nu_{\rm in}^2}.
\end{equation}
As before all quantities in the previous expressions have to be evaluated at $s=0$. We find a constant complex term, $\Lambda$, in the expansion of $\mathcal{D}$ (Equation~(\ref{eq:ds0})). The presence of this term removes the singularity of the solutions, so that $\mathcal{D} \ne 0$ for $s=0$. We use Equation~(\ref{eq:ds0}) to rewrite Equations~(\ref{eq:bas1s0}) and (\ref{eq:bas2s0}) as
\begin{eqnarray}
\frac{{\rm d} \xi_{{\rm i}r} }{{\rm d} s} &\approx& \frac{g_{\rm B}}{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{\Delta_{\rm A}} \frac{1}{\Lambda + s },\label{eq:bas1s02} \\
\frac{{\rm d} P}{{\rm d} s} &\approx& \frac{2}{\mu} \frac{B_\varphi B_z}{r_{\mathrm{A}}}\frac{f_{\rm B} }{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{\Delta_{\rm A}} \frac{1}{\Lambda + s }.\label{eq:bas2s02}
\end{eqnarray}
For our subsequent analysis it is convenient to use the scaled variable $\tau$ defined as
\begin{equation}
\tau = \frac{s + \Lambda_{{\rm R}}}{\Lambda_{{\rm I}}}.
\end{equation}
Then we integrate Equations~(\ref{eq:bas1s02}) and (\ref{eq:bas2s02}) and obtain
\begin{eqnarray}
\xi_{{\rm i}r} &\approx& \frac{g_{\rm B}}{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{\Delta_{\rm A}} \mathcal{G}(\tau) + \mathcal{C}_\xi, \\
P &\approx& \frac{2}{\mu} \frac{B_\varphi B_z}{r_{\mathrm{A}}}\frac{f_{\rm B} }{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{\Delta_{\rm A}} \mathcal{G}(\tau) + \mathcal{C}_P,
\end{eqnarray}
where $\mathcal{C}_\xi$ and $\mathcal{C}_P$ are constants of integration. The $\mathcal{G}(\tau)$ function is defined as
\begin{equation}
\mathcal{G}(\tau) \equiv \ln \left( \tau + i\right). \label{eq:psi}
\end{equation}
We separate the real and imaginary parts of $\mathcal{G}(\tau)$, namely
\begin{equation}
\Re \left( \mathcal{G}(\tau) \right)= \ln \sqrt{\tau^2 + 1}, \label{eq:psire}
\end{equation}
\begin{equation}
\Im \left( \mathcal{G}(\tau) \right)= \left\{
\begin{array}{lll}
\arctan \tau^{-1} + \pi, & \textrm{for} & \tau < 0, \\
\arctan \tau^{-1}, & \textrm{for} & \tau > 0.
\end{array} \right. \label{eq:psiim}
\end{equation}
Note that we need to add $\pi$ to the imaginary part of $\mathcal{G}(\tau)$ for $\tau < 0$ in order to make the solution continuous and analytic at $\tau=0$. This is not a problem since we can always incorporate a constant of integration.
From Equation~(\ref{eq:xiperp}) we obtain that the perpendicular displacement behaves as
\begin{equation}
\xi_{{\rm i}\perp} \approx \frac{\mathcal{C}_{\rm A}}{{\rho_{\rm i}} B } \frac{1}{\Delta_{\rm A} \Lambda_{{\rm I}}} \mathcal{F} (\tau),
\end{equation}
with $\mathcal{F}(\tau)$ defined as
\begin{equation}
\mathcal{F}(\tau) \equiv \frac{i}{\tau + i}. \label{eq:phi}
\end{equation}
We separate the real and imaginary parts of $\mathcal{F}(\tau)$, namely
\begin{equation}
\Re \left( \mathcal{F}(\tau) \right)= \frac{1}{\tau^2+1}, \qquad \Im \left( \mathcal{F}(\tau) \right)= \frac{\tau}{\tau^2+1}.
\end{equation}
Functions $\mathcal{F}(\tau)$ and $\mathcal{G}(\tau)$ play here the role of the {\em universal functions} found in a number of previous investigations of resonant waves in both stationary and non-stationary states and for different dissipative processes \citep[see, e.g.,][]{mok,goossens95,ruderman95,tirry,erdelyi,wrightallan,rudermanwright,vanlommel}. A comprehensive review on the importance and properties of the universal $\mathcal{F}(\tau)$ and $\mathcal{G}(\tau)$ functions can be found in \citet{goossens11}. In particular, our $\mathcal{F}(\tau)$ and $\mathcal{G}(\tau)$ functions coincide with the functions found by \citet{wrightallan} for magnetospheric Alfv\'en waves damped by Pedersen conductivity (see their Equations~(23) and (24)), and are also equivalent to the functions described by \citet{vanlommel} for non-stationary quasi-modes in the cusp continuum (see their Equation~(16)). Figure~\ref{fig:functions} displays the real and imaginary parts of functions $\mathcal{F}(\tau)$ and $\mathcal{G}(\tau)$.
\begin{figure}[!t]
\centering
\includegraphics[width=0.95\columnwidth]{18235f01a.eps}
\includegraphics[width=0.95\columnwidth]{18235f01b.eps}
\caption{Real (solid line) and imaginary (dashed line) of functions {\bf a)} $\mathcal{F}(\tau)$ and {\bf b)} $\mathcal{G}(\tau)$. \label{fig:functions}}
\end{figure}
Ion-neutral collisions generate a {\em collisional layer} around the location of the resonance, i.e., $\tau = 0$, in a similar way as resistivity and/or viscosity generate a dissipative layer \citep[see, e.g.,][]{hollwegyang,poedts,SGH91}. The thickness of the collisional layer, $\delta$, is proportional to $\Lambda_{{\rm I}}$, namely
\begin{equation}
\delta \sim \left| \Lambda_{{\rm I}} \right| = \frac{\omega^2}{\left| \Delta_{\rm A} \right|} \frac{\alpha \omega \nu_{\rm in}}{\omega^2 + \nu_{\rm in}^2}. \label{eq:thicknu}
\end{equation}
It can be seen that $\delta \to 0$ in both limits $\nu_{\rm in} \gg \omega$ and $\nu_{\rm in} \ll \omega$, whereas $\delta$ takes its maximum value for $\nu_{\rm in} = \omega$. This result points out that the collisional layer is extremely thin for realistic collision frequencies, i.e,, $\nu_{\rm in} \gg \omega$.
\subsection{Jump Conditions}
Here we use the expressions derived in the last section to determine the jump of the perturbations across the collisional layer. In the scaled variable $\tau$ and assuming $\omega > 0$, the limits $s \to \pm s_{\rm A}$ are equivalent to the limits $\tau \to \pm \textrm{sign} (\Delta_{\rm A}) \infty$. We denote by $[X]$ the jump of the quantity $X$ across the collisional layer, which we define as
\begin{equation}
[X] = \lim_{\tau\to \textrm{sign} (\Delta_{\rm A})\infty} X (\tau) - \lim_{\tau\to -\textrm{sign} (\Delta_{\rm A})\infty} X(\tau).
\end{equation}
Hence, the jumps of $\xi_{{\rm i}r}$ and $P$ are
\begin{eqnarray}
\left[\xi_{{\rm i}r}\right] &=& \frac{g_{\rm B}}{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{ \Delta_{\rm A} } [\mathcal{G}], \\
\left[ P \right] &=& \frac{2}{\mu} \frac{B_\varphi B_z}{r_{\mathrm{A}}}\frac{f_{\rm B} }{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{\Delta_{\rm A} }[\mathcal{G}].
\end{eqnarray}
To calculate the jumps of $\xi_{{\rm i}r}$ and $P$ we need the jump of $\mathcal{G}$, namely
\begin{equation}
\left[\mathcal{G}\right] = - i \textrm{sign} (\Delta_{\rm A}) \pi.
\end{equation}
Then it is straightforward to obtain
\begin{eqnarray}
[\xi_{{\rm i}r}] &=& - i \pi \frac{g_{\rm B}}{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{\left| \Delta_{\rm A} \right|} , \label{eq:jumpxi} \\
\left[ P \right] &=& - i \pi \frac{2}{\mu} \frac{B_\varphi B_z}{r_{\mathrm{A}}}\frac{f_{\rm B} }{\mu {\rho_{\rm i}}^2 v_{\mathrm{Ai}}^2} \frac{\mathcal{C}_{\rm A}}{\left| \Delta_{\rm A} \right|}. \label{eq:jumpp}
\end{eqnarray}
We find the remarkable result that these jumps are the same as obtained for ideal and dissipative MHD \citep[e.g.,][]{SGH91,goossens95,ruderman95}. This means that ion-neutral collisions do not modify the jumps of the plasma perturbations across the resonant layer. In the particular case of a straight magnetic field, i.e., $B_\varphi = 0$, the conserved quantity across the resonant layer is proportional to the total pressure perturbation, and the jump conditions for $\xi_{{\rm i}r}$ and $P$ (Equations~(\ref{eq:jumpxi}) and (\ref{eq:jumpp})) become
\begin{equation}
[\xi_{{\rm i}r}] = - i \pi \frac{m^2/r_{\mathrm{A}}^2}{{\rho_{\rm i}} \left| \Delta_{\rm A} \right|} P, \qquad \left[ P \right] = 0. \label{eq:jump0}
\end{equation}
Finally, the asymptotic expansion of function $\mathcal{F}$ for $\tau \to \pm \infty$ gives as the asymptotic behavior of $\xi_{{\rm i}\perp}$ when we move away from the resonance position as
\begin{equation}
\xi_{{\rm i}\perp} \sim \frac{\mathcal{C}_{\rm A}}{{\rho_{\rm i}} B} \frac{1}{\Delta_{\rm A} \Lambda_{{\rm I}}} \frac{i}{ \tau}, \label{eq:jumpxip}
\end{equation}
which again corresponds with the asymptotic dependence $\tau^{-1}$ found in both ideal and dissipative MHD.
In this Section we have studied the case of propagating waves, i.e., real $\omega$ and complex $k_z$. However, most of the analysis is equivalent for the case of standing quasi-modes, i.e., complex $\omega$ and real $k_z$. In particular, the jump conditions (Equations~(\ref{eq:jumpxi}) and (\ref{eq:jumpp})) are exactly the same for both propagating and standing modes. We refer the reader to, e.g., \citet{tirry} and \citet{vanlommel} for details about the derivation of the jump conditions for standing quasi-modes.
\section{Kink waves in straight tubes}
\label{sec:tube}
Here we perform an application of the theory of Section~\ref{sec:resonant} to the case of propagating resonant kink waves in straight flux tubes.
We consider a straight magnetic cylinder of radius $R$ with a constant vertical magnetic field, i.e., $B_\varphi = 0$ and $B_z =$~constant. The tube is homogeneous in the longitudinal direction. In the radial direction, the tube is composed of an internal region with constant ion density ${\rho_{\rm i}}_{\rm 1}$, a nonuniform transitional layer of thickness $l$, and an external region with constant ion density ${\rho_{\rm i}}_{\rm 2}$. In the nonuniform layer the density changes continuously from the internal to the external values in the interval $R-l/2 < r < R+l/2$. Hereafter, indices `1' and `2' refer to the internal and external regions, respectively. For the sake of simplicity, the parameter $\alpha$ is taken constant everywhere so that the neutral density, ${\rho_{\rm n}}$, follows the radial dependence of the ion density. The ion-neutral collision frequency, $\nu_{\rm in}$, is also a constant to simplify matters.
We restrict ourselves to the case of thin tubes, i.e., the wavelength is much longer than the radius of the tube. For propagating waves the thin tube (TT) approximation is equivalent to the low-frequency approximation, i.e., $\omega R / v_{\rm k} \ll 1$, with $v_{\rm k}$ the kink velocity of ions defined as
\begin{equation}
v_{\rm k} = \sqrt{\frac{{\rho_{\rm i}}_{\rm 1} {v^2_{\rm Ai}}_{\rm 1} + {\rho_{\rm i}}_{\rm 2} {v^2_{\rm Ai}}_{\rm 2}}{{\rho_{\rm i}}_{\rm 1} + {\rho_{\rm i}}_{\rm 2}}}.
\end{equation}
To check whether this approximation is realistic, let us consider the properties of the observed waves propagating in chromospheric spicules \citep[e.g.,][]{depontieu07, okamotodepontieu}. Taking $R=200$~km for the spicule radius and $v_{\rm k} = 270$~km~s$^{-1}$ for the averaged phase velocity as estimated by \citet{okamotodepontieu}, we obtain $\omega R / v_{\rm k} \approx$~0.1 for a period of 45~s \citep{okamotodepontieu} and $\omega R / v_{\rm k} \approx$~0.02 for a period of 3~min \citep{depontieu07}. Therefore, the use of the TT approximation is justified.
Additionally, we use the Thin Boundary (TB) approximation \citep[see, e.g.,][]{hollwegyang} and assume $l/R \ll 1$. The TB approximation enables us to use the jump conditions given in Equation~(\ref{eq:jump0}) as boundary conditions for the wave perturbations at the tube boundary. Detailed explanations of this method can be found in, e.g., \citet{goossens06} and \citet{goossens08}. Doing so, we arrive at the dispersion relation for resonant kink ($m=1$) and fluting ($m \geq 2$) waves in the TT and TB approximations. For simplicity we omit the intermediate steps which can be found in, e.g., \citet{goossens92}. The dispersion relation is
\begin{equation}
\frac{1}{{\rho_{\rm i}}_{\rm 1} \left( \tilde{\omega}^2 - {\omega_{\rm A}^2}_{\rm 1} \right)} + \frac{1}{{\rho_{\rm i}}_{\rm 2} \left( \tilde{\omega}^2 - {\omega_{\rm A}^2}_{\rm 2} \right)} = i \pi \frac{m/R}{{\rho_{\rm i}}(R) \left| \Delta_{\rm A} \right|_R}. \label{eq:disperthin}
\end{equation}
In Equation~(\ref{eq:disperthin}) we approximated $r_{\mathrm{A}} \approx R$. The effect of ion-neutral collisions is enclosed in the definition of $\tilde{\omega}$ (Equation~(\ref{eq:omegat})). For fixed and real $\omega$ the solution of Equation~(\ref{eq:disperthin}) is a complex longitudinal wavenumber, $k_z = k_{z \rm R} + i k_{z \rm I}$, where $k_{z \rm R}$ and $k_{z \rm I}$ are the real and imaginary parts of $k_z$, respectively. In the present normal mode analysis, which represents the stationary state of wave propagation, the amplitude of the propagating wave is proportional to $\exp \left( - z/L_{\rm D} \right)$, with $L_{\rm D}$ the damping length defined as
\begin{equation}
L_{\rm D} = \frac{1}{k_{z \rm I}}.
\end{equation}
We investigate kink waves and set $m=1$. In the absence of resonant absorption the right-hand side of Equation~(\ref{eq:disperthin}) is zero. In such a case, we can obtain an approximate expression for $k_{z \rm R}^2$ by assuming $k_{z \rm I}^2 \ll k_{z \rm R}^2$, namely
\begin{equation}
k_{z \rm R}^2 \approx \frac{\omega^2}{v_{\rm k}^2} \frac{\omega^2+ \left( 1+\alpha \right) \nu_{\rm in}^2}{\omega^2+ \nu_{\rm in}^2}. \label{eq:kzr}
\end{equation}
In the single-fluid limit, $\nu_{\rm in} \gg \omega$ and $k_{z \rm R}^2 \approx \omega^2 \left( 1+\alpha \right) / v_{\rm k}^2$, while in the fully ionized case $\alpha=0$ and $k_{z \rm R}^2 \approx \omega^2 / v_{\rm k}^2$. Note that the expressions for $k_{z \rm R}^2$ in the single-fluid and fully ionized cases differ by a factor $\left( 1+\alpha \right)$. SOB studied propagating kink waves in the single-fluid approximation and found the same expression for $k_{z \rm R}^2$ for both partially ionized and fully ionized plasmas (see their Equations~(11) and (15)). The reason is that SOB used a slightly different definition of the kink velocity, $v_{\rm k}$. Here, the kink velocity is defined using the ion density only, while SOB defined $v_{\rm k}$ using the total (ion + neutral) density. In SOB the total density is fixed and the effect of the ionization degree is to change the relative ion and neutral densities. On the contrary, here the ion density is a fixed parameter while the amount of neutrals depends on the ionization degree.
\subsection{Approximate expression for the damping length}
Here we seek an approximate expression for $L_{\rm D}$. First we evaluate the factor ${\rho_{\rm i}}(R) \left| \Delta_{\rm A} \right|_R$ in Equation~(\ref{eq:disperthin}) by using the resonant condition $\omega^2 = \omega_{\rm A}^2 \left( R \right)$, namely
\begin{equation}
{\rho_{\rm i}}(R) \left| \Delta_{\rm A} \right|_R = \omega^2 \left| \frac{{\rm d}{\rho_{\rm i}}}{{\rm d}r} \right|_R,
\end{equation}
where $\left| {\rm d}{\rho_{\rm i}} / {\rm d}r \right|_R$ is the radial derivative of the density profile at $r = R$. Next we write $k_z = k_{z \rm R} + i k_{z \rm I}$ and assume weak damping, so we neglect terms with $\mathcal{O} \left( k_{z \rm I}^2 \right)$. An expression for the ratio $k_{z \rm I} / k_{z \rm R}$ is obtained from Equation~(\ref{eq:disperthin}), namely
\begin{equation}
\frac{k_{z \rm I}}{k_{z \rm R}} \approx \frac{1}{2} \frac{\alpha \omega \nu_{\rm in}}{\omega^2 + \left( 1 + \alpha \right) \nu_{\rm in}^2} + \frac{\pi}{8} \frac{1}{R} \frac{\left( {\rho_{\rm i}}_{\rm 1} - {\rho_{\rm i}}_{\rm 2} \right)^2}{{\rho_{\rm i}}_{\rm 1} + {\rho_{\rm i}}_{\rm 2}} \frac{1}{\left| {\rm d}{\rho_{\rm i}} / {\rm d}r \right|_R}. \label{eq:ratiokz}
\end{equation}
Now we express $\left| {\rm d}{\rho_{\rm i}} / {\rm d}r \right|_R$ as
\begin{equation}
\left| \frac{{\rm d}{\rho_{\rm i}}}{{\rm d}r} \right|_R = F \frac{\pi^2}{4} \frac{{\rho_{\rm i}}_{\rm 1} - {\rho_{\rm i}}_{\rm 2}}{l},
\end{equation}
with $F$ a numerical factor that depends on the form of the density profile. For example, $F = 4/\pi^2$ for a linear profile \citep{goossens2002} and $F = 2/\pi$ for a sinusoidal profile \citep{rudermanroberts}. We assume that $k_{z \rm R}$ is approximately the same as in case without resonant damping (Equation~(\ref{eq:kzr})) and work on Equation~(\ref{eq:ratiokz}) to find the expression for $L_{\rm D}$ as,
\begin{equation}
\frac{1}{L_{\rm D}} \approx \frac{1}{L_{\rm D, RA}} + \frac{1}{L_{\rm D, IN}}, \label{eq:ldgen}
\end{equation}
with $L_{\rm D, RA}$ and $L_{\rm D, IN}$ the damping lengths due to resonant absorption and ion-neutral collisions, respectively, given by
\begin{eqnarray}
L_{\rm D, RA} &=& 2 \pi \mathcal{F} v_{\rm k} \frac{R}{l} \frac{\zeta + 1}{\zeta - 1} \frac{1}{\omega} \left( \frac{\omega^2+\nu_{\rm in}^2}{\omega^2+ \left( 1+\alpha \right) \nu_{\rm in}^2} \right)^{1/2}, \label{eq:ldra}\\
L_{\rm D, IN} &=& 2 v_{\rm k} \frac{\left( \omega^2 + \left( 1 + \alpha \right) \nu_{\rm in}^2 \right)^{1/2} \left( \omega^2+\nu_{\rm in}^2 \right)^{1/2}}{\alpha \omega^2\nu_{\rm in}} , \label{eq:ldin}
\end{eqnarray}
with $\zeta = {\rho_{\rm i}}_1 / {\rho_{\rm i}}_2$ the ion density contrast. Importantly, we find that both the collision frequency, $\nu_{\rm in}$, and the ionization degree, $\alpha$, are present in the expression of the damping length by resonant absorption (Equation~(\ref{eq:ldra})). In the single-fluid limit ($\nu_{\rm in} \gg \omega$) Equations~(\ref{eq:ldra}) and (\ref{eq:ldin}) become
\begin{eqnarray}
L_{\rm D, RA} &\approx & 2 \pi \mathcal{F} v_{\rm k} \frac{R}{l} \frac{\zeta + 1}{\zeta - 1} \frac{1}{\omega} \left( 1+\alpha \right)^{-1/2}, \label{eq:ldrasf}\\
L_{\rm D, IN} &\approx & 2 v_{\rm k} \frac{\left( 1+\alpha \right)^{1/2} \nu_{\rm in}}{\alpha} \frac{1}{\omega^2}. \label{eq:ldinsf}\
\end{eqnarray}
The damping length due to resonant absorption in the single-fluid limit is inversely proportional to the frequency as in TGV. Indeed, in the fully ionized case ($\alpha=0$) Equations~(\ref{eq:ldra}) and (\ref{eq:ldrasf}) become Equation~(22) of TGV. For a partially ionized plasma $L_{\rm D, RA}$ also depends on the ionization degree, $\alpha$. The dependence on the ionization degree was not discussed by SOB. SOB obtained an expression for the ratio of the damping length to the wavelength (see their Equation~(20)) and did not explicitly write the expression for the damping length. It turns out that the factor containing the ionization degree cancels out when the ratio of the damping length to the wavelength is computed, and so the dependence of $L_{\rm D, RA}$ on the ionization degree was not noticed by SOB. The damping length by ion-neutral collisions in the single-fluid limit (Equation~(\ref{eq:ldinsf})) is inversely proportional to $\omega^{2}$. This result is consistent with the expressions found by, e.g, \citet{hearendel,pecseli,depontieu98}, and SOB among others.
Figure~\ref{fig:ld} shows $L_{\rm D}/R$ versus $\nu_{\rm in}/\omega$ for a particular set of parameters given in the caption of the Figure. The total damping length (solid line) has a minimum for $\nu_{\rm in} \sim \omega$. The reason for this behavior is that damping by ion-neutral collisions (dotted line) becomes more relevant than resonant damping (dashed line) for $\nu_{\rm in} \sim \omega$. In both the collisionless ($\nu_{\rm in} \ll \omega$) and single-fluid ($\nu_{\rm in} \gg \omega$) limits, the total damping length is well approximated by the damping length due to resonant absorption. In particular, for $\nu_{\rm in} \gg \omega$ the result tends to that predicted by Equation~(\ref{eq:ldrasf}).
\begin{figure}[!t]
\centering
\includegraphics[width=0.95\columnwidth]{18235f02.eps}
\caption{$L_{\rm D}/R$ vs. $\nu_{\rm in}/\omega$ for the kink wave in a magnetic flux tube with $l/R=0.2$, $\alpha=0.5$, $\zeta=10$, $F=2/\pi$, and $\omega R / v_{\rm k} = 0.1$. The solid line is the total damping length in the TT and TB approximations computed from Equation~(\ref{eq:ldgen}). The dashed line is the damping length due to resonant absorption (Equation~(\ref{eq:ldra})) only, and the dotted line is the damping length due to ion-neutral collisions (Equation~(\ref{eq:ldin})) only. Symbols $\Diamond$ correspond to the full numerical eigenvalue results. \label{fig:ld}}
\end{figure}
Now we fix the collision frequency to $\nu_{\rm in} R / v_{\rm k} = 100$ and compute $L_{\rm D}/R$ as a function of $\omega R / v_{\rm k} $ (see Figure~\ref{fig:ldw}). The remaining parameters are the same as in Figure~\ref{fig:ld}. Although the observed wave frequencies in the solar atmosphere correspond to $\omega R / v_{\rm k} \ll 1$, it is interesting to perform a general study for high values of the frequency. Our results point out that the damping of the kink wave is governed by different mechanisms depending on the value of $\omega R / v_{\rm k} $. We find that the damping length for $\omega R / v_{\rm k} \ll 1$ is dominated by resonant absorption. Ion-neutral collisions start to become important when $\omega R / v_{\rm k} \sim 10$. Finally, collisions are the dominant damping mechanism for large $\omega R / v_{\rm k}$. We must note that the analytical solution in the TT approximation may not provide accurate results for values of $\omega R / v_{\rm k}$ departing from the limit $\omega R / v_{\rm k} \ll 1$. For this reason, numerical eigenvalue computations which overcome the limitations of the analytical approximations are performed in the next Subsection.
\begin{figure}[!t]
\centering
\includegraphics[width=0.95\columnwidth]{18235f03.eps}
\caption{$L_{\rm D}/R$ vs. $\omega R / v_{\rm k} $ with $\nu_{\rm in} R / v_{\rm k} = 100$. The remaining parameters and the meaning of the line styles are the same as in Figure~\ref{fig:ld}. \label{fig:ldw}}
\end{figure}
\subsection{Numerical results}
Here we compare the approximate analytical result obtained by solving the dispersion relation in the TT and TB approximations (Equation~(\ref{eq:ldgen})) with the damping length obtained by solving the full eigenvalue problem numerically. The numerical solution is not limited by the TT and TB approximations. The numerical code is similar to that used in TGV and \citet{resonantflow}. The reader is refereed to these papers for details of the numerical scheme. In short, Equations~(\ref{eq:momion})--(\ref{eq:induct}) are integrated in the radial direction assuming a time dependence of the form $\exp \left( - i \omega t \right)$ and a spatial dependence in the $\varphi$ and $z$ directions as $\exp \left( i \varphi + i k_z z \right)$. The code solves the eigenvalue problem for the temporal damping of standing waves, i.e., complex $\omega$ provided a fixed and real $k_z$. To study spatial damping we need to convert the results from complex $\omega$ and real $k_z$ to real $\omega$ and complex $k_z$. The conversion is done following the method explained in TGV (see their Equation~(40)).
First, we consider the same parameters as in Figure~\ref{fig:ld} and compute the numerically determined $L_{\rm D}/R$ versus $\nu_{\rm in}/\omega$. To compare with the analytical approximation, the numerical result is plotted using symbols $\Diamond$ in Figure~\ref{fig:ld}. A very good agreement between approximate and numerical results is found. This means that for $\omega R / v_{\rm k} \ll 1$ the approximate analytical theory provides accurate results.
Next we numerically explore the effect of increasing the wave frequency. We take the same parameters as in Figure~\ref{fig:ldw}. Again, we use symbols $\Diamond$ to represent the eigenvalue result in Figure~\ref{fig:ldw}. We find that in the eigenvalue computations the transition between the regime dominated by resonant damping and that dominated by collisional damping occurs around $\omega R / v_{\rm k} \sim 1$, while in the analytical approximation collisions start to become important for $\omega R / v_{\rm k} \sim 10$. This discrepancy is an effect of the TT approximation. For realistic values of the wave frequency ($\omega R / v_{\rm k} \sim 10^{-2} - 10^{-1}$) both numerical and analytic results are in excellent agreement.
\section{Chromospheric kink waves}
\label{sec:chromos}
The results of Section~\ref{sec:tube} have direct implications for MHD waves propagating in partially ionized plasmas of the solar atmosphere. For kink waves studied in this paper, both resonant absorption and ion-neutral collisions decrease the amplitude of the waves. However, the two processes represent very different physical mechanisms.
On the one hand, resonant absorption is an ideal process that transfers wave energy from global kink motions to localized azimuthal motions within the transversely inhomogeneous part of the flux tube. These azimuthal motions keep propagating along magnetic field lines \citep[see the numerical simulations by, e.g.,][]{pascoe1,pascoe2}. A detailed investigation of the energy transfer in the case of standing waves was done in \citet{arregui2d} by analyzing the Poynting flux in a two-dimensional configuration. However, resonant absorption itself does not dissipate wave energy in the plasma. The energy fed into the inhomogeneous layer will be dissipated by another mechanism later \citep[see, e.g., the results of][ in resistive MHD]{poedtskerner, poedtskerner2,poedts,poedts2,poedts3}. Hence, the damping length due to resonant absorption, $L_{\rm D, RA}$, represents the length scale for the kink motions to be converted into azimuthal motions.
On the other hand, ion-neutral collisions is a true dissipative process which deposits wave energy {\em in situ} and so it contributes to plasma heating. Hence, the damping length due to ion-neutral collisions, $L_{\rm D, IN}$, represents the length scale for the kink wave energy to be dissipated by ion-neutral collisions. By comparing the values of $L_{\rm D, RA}$ and $L_{\rm D, IN}$ we can estimate the fraction of energy converted to Alfv\'enic, azimuthal motions and the fraction of energy dissipated by collisions.
Let us apply this theory to kink waves propagating along chromospheric waveguides (spicules). We assume that the driver of the waves is located at the photosperic level and the waves propagate through the chromosphere to the corona. We take the variation of physical parameters with height (e.g., density, temperature, ionization degree, etc.) from the VALC model \citep{valc}. For the chromospheric magnetic field we consider the model used by \citet{leakearber}. Then we use Equations~(\ref{eq:ldrasf}) and (\ref{eq:ldinsf}) to compute the values of $L_{\rm D, RA}$ and $L_{\rm D, IN}$. As the physical parameters change along the spicule, both $L_{\rm D, RA}$ and $L_{\rm D, IN}$ are functions of height in the chromosphere. We plot in Figure~\ref{fig:averagedld}(a) the values of the damping lengths as functions of height for a wave period of 45~s \citep{okamotodepontieu}. The damping length due to resonant absorption increases with height. This is an effect of the increase of the kink velocity, $v_{\rm k}$, with height \citep{stratified}. At low heights, $L_{\rm D, RA}$ is comparable to the thickness of the whole chromosphere, meaning that a large fraction of wave energy is in the form of azimuthal motions when the wave reaches the coronal level. On the contrary, $L_{\rm D, IN}$ is several orders of magnitude longer. As $L_{\rm D, IN}$ decreases with height, the effect of collisions is more important in the upper chromosphere.
\begin{figure}[!t]
\centering
\includegraphics[width=0.95\columnwidth]{18235f04a.eps}
\includegraphics[width=0.95\columnwidth]{18235f04b.eps}
\caption{(a) $L_{\rm D, RA}$ (solid line) and $L_{\rm D, IN}$ (dashed line) vs. height above the photosphere for a wave period of 45~s. (b) Averaged $L_{\rm D, RA}$ (solid line) and $L_{\rm D, IN}$ (dashed line) in the chromosphere vs. wave period. The horizontal dotted line represents the height of the chromosphere above the photosphere. \label{fig:averagedld}}
\end{figure}
Next, we calculate the damping length averaged along the spicule, $\bar{L}_{\rm D}$, as
\begin{equation}
\bar{L}_{\rm D} = \frac{1}{H} \int_0^H L_{\rm D} (s) {\rm d} s, \label{eq:averagedld}
\end{equation}
where $s$ represents the direction along the spicule and $H$ is the height of the chromosphere above the photosphere. We take $H=$~3,000~km. Equation~(\ref{eq:averagedld}) is used to calculate the averaged values of both $L_{\rm D, RA}$ and $L_{\rm D, IN}$. We plot in Figure~\ref{fig:averagedld}(b) the averaged values of the damping lengths as functions of the wave period. First, we obtain that the averaged $L_{\rm D, IN}$ is several orders of magnitude longer than the averaged $L_{\rm D, RA}$ in the range of periods taken into account in Figure~\ref{fig:averagedld}(b). This means that ion-neutral collisions have little impact on wave propagation. On the contrary, the averaged $L_{\rm D, RA}$ for periods less than 10~s is smaller than or of the same order as the height of the chromosphere. This result points out that only waves with periods longer than 10~s are able to reach the coronal level in the form of kink motions. Waves with shorter periods reach the corona as small-scale azimuthal motions, which are unobservable with present day instruments. This effectively imposes a lower limit for the period of kink waves observable in the corona. This is consistent with the observed periods of coronal waves \citep[e.g.,][]{tomczyk07,tomczyk09,mcintosh2011}.
\section{Conclusion}
\label{sec:con}
In this paper we have investigated resonant Alfv\'en waves in partially ionized plasmas. We find that the conserved quantity at the resonance and the jump of the perturbations across the resonant layer are the same as in fully ionized, ideal plasmas. We have derived expressions for the damping lengths due to resonant absorption and due to ion-neutral collisions for the case of propagating kink waves in straight magnetic tubes. In the limit of large collision frequencies, the damping length due to resonant absorption is inverselly proportional to the frequency as in the fully ionized case \citep[see][]{TGV}, whereas the damping length due to ion-neutral collisions is inverselly proportional to the square of the frequency. We have applied the theory to the case of chromospheric kink waves propagating from the photosphere to the corona. We conclude that the solar chromosphere acts as a filter for kink waves. Waves with peridos shorter than 10~s, approximetely, reach the corona in the form of small-scale azimuthal waves. Only waves with periods longer than 10~s can be observed in the corona as kink waves.
\acknowledgements{
We thank I. Arregui and A. J. D\'iaz for useful comments. R.S. acknowledges support from a Marie Curie Intra-European Fellowship within the European Commission 7th Framework Program (PIEF-GA-2010-274716). M.G. acknowledges support from K.U. Leuven via GOA/2009-009}
|
\section{Introduction}
On March 11, 2011, reactors at the Fukushima Dai-ichi nuclear power plant in Japan (37\degree45' N, 141\degree27' E) were shut down following a 9.0 magnitude earthquake. Emergency diesel generators were activated to power water pumps needed to cool the reactors and prevent intensely radioactive nuclear fuel from overheating and damaging the reactor containment vessels. Shortly after the earthquake, the plant was struck by a 14-m tsunami which flooded the electrical building, disabling the emergency generators~\citep{wnn}. Fires, explosions and possible partial core meltdowns released radioactive fission products into the atmosphere. On April 20, 2011, the International Atomic Energy Agency (IAEA) reported \textsuperscript{131}I deposition in Japan ranging from 1.8 to 368 Bq/m\textsuperscript{2} in 13 prefectures and \textsuperscript{137}Cs deposition ranging from 2.4 to 160 Bq/m\textsuperscript{2} in seven prefectures~\citep{iaea}.
Uncertainty surrounded the development of the situation at Fukushima as plant staff struggled to restore power and to adequately cool the reactors. A limited amount of information and measurements relevant to the release of radioactive material had been available. Considering global interest in possible public health effects and local impact on our physics program that relies on ultra-low radioactive background detectors, we began collecting airborne particle samples and monitoring for fission products on March 17, 2011.
Airborne fission products, including \textsuperscript{131,132}I, \textsuperscript{134,137}Cs, \textsuperscript{132}Te and \textsuperscript{133}Xe were first detected on the west coast of the United States on March 16, 2011, and reached a maximum activity concentration of 4.4 $\pm$ 1.3 mBq/m\textsuperscript{3} of \textsuperscript{131}I 3 d after the first detection of fission products~\citep{Bow11,Leo11}. Radioactivity has also been measured in Europe~\citep{Mas11}, western Japan~\citep{Fus11}, Russia~\citep{Bol11}, in rainwater in the United States~\citep{Nor11} and is continuously monitored in France~\citep{IRSN11}.
We first detected airborne fission products in Chapel Hill between 20:00 UTC on March 18, 2011, and 20:00 UTC on March 19, 2011. We measured a maximum activity concentration of 4.2 $\pm$ 0.6 mBq/m\textsuperscript{3} of \textsuperscript{131}I in the interval between March 29, 2011 and March 30, 2011. The time dependence of \textsuperscript{131}I and \textsuperscript{137}Cs activity seems to have been dominated by local rain, but we are able to draw some conclusions about the nature of the reactor accident and the transport of fission products around the world.
\section{Materials and Methods}
The air sampling pump and filter were housed on the roof of Phillips Hall at the University of North Carolina at Chapel Hill (35\degree55'N, 79\degree2' W, 150 m elevation). The pump is a Staplex model TFIA 110-125 V DC/AC with built-in flow meter and an 8'' $\times$ 10'' filter holder assembly \citep{samplers}. With a filter in place, the flow rate of the air sampler was measured to be 2050 m\textsuperscript{3}/d. Air filters collected airborne particles for approximately 24 h, then were removed, folded into 5 cm $\times$ 10 cm rectangles and sealed in nylon bags. A total of 38 air filters were collected over 62 d following the March 11, 2011 earthquake.
Two 5\% relative efficiency to NaI (R.E.) high-purity germanium (HPGe) detectors were used to identify characteristic gamma rays from radioactive fission products. The detectors were setup horizontally with a 10 mm separation between their end caps. Bagged filters were placed between the end caps of the detectors for assay. The detectors were surrounded by a lead shield of 5-10 cm thickness to reduce backgrounds from cosmic rays and natural radiation.
Ten filters which had already been measured using the setup described above were assayed at the Kimballton Underground Research Facility (KURF) in order to search for more exotic, long-lived fission products. KURF is located in the Kimballton mine near Ripplemead, VA, which provides shielding equivalent to 1450 m of water from cosmic-ray induced backgrounds. The detectors in this counting facility have a background rate about 40 times less than the above-ground detectors in the region of 40 -- 2700 keV. All 10 filters were tightly fit into a Marinelli beaker and were assayed simultaneously in the `VT-1' detector which is a 35\% R.E. coaxial HPGe detector with very low background (ORTEC LLB). For technical specifications on the counting systems and facility see~\citet{Fin10}.
A single filter was also measured using the VT-1 detector to determine the gamma-ray detection efficiency of the surface based counting system. The efficiency of the VT-1 detector had previously been determined using an existing Monte Carlo simulation that had been calibrated using point sources of known activity spanning wide gamma-ray energy range in multiple locations surrounding the detector. The detection efficiencies of the above-ground detectors were determined by comparing \textsuperscript{131}I and \textsuperscript{137}Cs count rates to VT-1 measurements of the same filter. The relative efficiencies were also verified by comparing the efficiency of a \textsuperscript{133}Ba (E$_\gamma$ = 356 keV) point source on the top of the cryostat of the VT-1 detector to a similar source of known activity in the same position on the cryostats of the above-ground detectors.
To determine the filter efficiency for air particles containing \textsuperscript{131}I and \textsuperscript{137}Cs, two filters were placed in the air sampler in series and the activities of the two filters were compared. Using this method, the filter efficiency for particles carrying \textsuperscript{131}I was determined to be 87\% $\pm$ 5\%. The filter efficiency for air particles containing \textsuperscript{137}Cs was determined to be 98.7\% $\pm$ 0.6\%. The manufacturer quotes a filter efficiency of 99.98\% for particles down to 0.3 $\mu$m \citep{filters}. It should be noted that the measured \textsuperscript{131}I only represents particulate species collected in the air filters. This accounts for only about 50\% of the total \textsuperscript{131}I in the air. The rest is distributed in gases such as I$_2$ and methyl iodide \citep{Per90}.
The particulate activity concentration from a given isotope can be calculated as:
\begin{equation}
A_{air} = \frac{R}{\epsilon_{det}\epsilon_{filter}}\frac{1}{Qt_{samp}}
\end{equation}
\noindent where $\epsilon_{det}$ is the absolute gamma-ray detection efficiency, $\epsilon_{filter}$ is the air filter efficiency, $Q$ is the flow rate of the air sampler, t$_{samp}$ is the time duration of air sampling and $R$ is the count rate of the filter at the end of the sampling period. $R$ has been corrected for any delay between sampling and counting using the known lifetimes of the relevant isotopes. We did not correct for decays during the sampling period.
\section{Results and Discussion}
The first filter was placed in the air sampler on March 17, 2011, and collected particles for 24 h. Although there are many nuclear facilities in the United States including the Shearon Harris nuclear power plant about 30 km from the sampling site~\citep{nrc2}, no fission products were detected in this filter and all visible gamma-ray peaks could be attributed to known background sources. Filters from subsequent days produced clear peaks attributable to \textsuperscript{131}I and \textsuperscript{137}Cs as seen in Fig.~\ref{fig:arrival}. The sampling dates and measured air activity concentrations for the first 62 d of data are listed in Table \ref{samplingdata}. A plot of the \textsuperscript{131}I and \textsuperscript{137}Cs activity concentrations measured over the first 40 d is shown in Fig.~\ref{time}.
\begin{figure*} [htp]
\centering
\subfloat[]{\epsfig{file=Fig1a.eps, width=.45\textwidth}}
\subfloat[]{\epsfig{file=Fig1b.eps, width=.45\textwidth}}\\
\caption{Comparison of gamma-ray spectra from the filter sampled from March 18 - 19, 2011 (gray) and March 20 - 21, 2011 (black) showing an increase in fission fragments. (a) The dominant fission fragment peak is at 364 keV from the decay of \textsuperscript{131}I. The peak at 352 keV is from \textsuperscript{214}Pb, which comes from \textsuperscript{222}Rn in the air. (b) Fission fragment peak from \textsuperscript{137}Cs at 662 keV.}
\label{fig:arrival}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=.9\textwidth, angle=0]{Fig2.eps}
\caption{The time dependence of \textsuperscript{131}I and \textsuperscript{137}Cs activity and local rainfall. Time is shown in days after the March 11 earthquake (rounded to the nearest day). Each activity measurement corresponds to the day the filter was removed from the air sampler. Days when the fission product activity was below the detection sensitivity are reported as upper limits (95\% confidence limit). The data may also be found in Table \ref{samplingdata}.}
\label{time}
\end{figure*}
\begin{table*}
\small
\caption{\ Measured \textsuperscript{131}I and \textsuperscript{137}Cs air activity concentrations during the first 62 d of air sampling. Filters collected air particles for approximately 24 h.}
\label{samplingdata}
\begin{tabular*}{.99\textwidth}{@{\extracolsep{\fill}}lllll}
\hline
\\
Filter sampling & Days since & Precipitation\textsuperscript{b} & \textsuperscript{131}I activity & \textsuperscript{137}Cs activity\\
end date\textsuperscript{a} & earthquake & (cm) & concentration & concentration \\
& & & (mBq/m\textsuperscript{3}) & (mBq/m\textsuperscript{3}) \\
\hline
March 19 &9 & 0.00 & 0.36 $\pm$ 0.07 & 0.11 $\pm$ 0.03\\
March 21 &11 & 0.00 & 0.55 $\pm$ 0.09 & 0.13 $\pm$ 0.03\\
March 22 &12 & 0.00 & 0.52 $\pm$ 0.09 & 0.16 $\pm$ 0.04\\
March 24 &14 & 0.38 & 0.67 $\pm$ 0.11 & 0.06 $\pm$ 0.02\\
March 25 &15 & 0.00 & 0.85 $\pm$ 0.13 & 0.11 $\pm$ 0.03\\
March 26 &16 & 0.44 & 2.42 $\pm$ 0.37 & 0.18 $\pm$ 0.04\\
March 27 &17 & 0.18 & 0.88 $\pm$ 0.14 & 0.08 $\pm$ 0.02\\
March 29 &19 & 0.13 & 3.40 $\pm$ 0.51 & 0.18 $\pm$ 0.04\\
March 30 &20 & 2.44 & 4.15 $\pm$ 0.62 & 0.42 $\pm$ 0.07\\
March 31 &21 & 0.03 & 0.93 $\pm$ 0.15 & 0.09 $\pm$ 0.03\\
April 1 &22 & 0.00 & 2.66 $\pm$ 0.40 & 0.21 $\pm$ 0.04\\
April 2 &23 & 0.00 & 3.99 $\pm$ 0.60 & 0.33 $\pm$ 0.06\\
April 3 &24 & 0.00 & 1.24 $\pm$ 0.19 & 0.06 $\pm$ 0.02\\
April 4 &25 & 0.00 & 0.92 $\pm$ 0.14 & 0.09 $\pm$ 0.03\\
April 5 &26 & 0.43 & 0.23 $\pm$ 0.05 & 0.08 $\pm$ 0.02\\
April 6 &27 & 0.00 & 0.83 $\pm$ 0.13 & 0.10 $\pm$ 0.03\\
April 7 &28 & 0.00 & 0.66 $\pm$ 0.11 & 0.08 $\pm$ 0.03\\
April 8 &29 & 0.33 & 0.48 $\pm$ 0.08 & 0.09 $\pm$ 0.02\\
April 9 &30 & 2.87 & 0.26 $\pm$ 0.05 & 0.12$\pm$ 0.04\\
April 10 &31 & 0.00 & $<$ 0.06 & 0.08 $\pm$ 0.03\\
April 11 &32 & 0.00 & $<$ 0.06 & $<$ 0.04\\
April 12 &33 & 0.41 & 0.12 $\pm$ 0.03 & 0.09 $\pm$ 0.04\\
April 13 &34 & 0.00 & $<$ 0.05 & $<$ 0.04\\
April 14 &35 & 0.00 & 0.15 $\pm$ 0.04 & $<$ 0.04\\
April 15 &36 & 0.00 & 0.16 $\pm$ 0.04 & 0.10 $\pm$ 0.02\\
April 16 &37 & 1.52 & 0.12 $\pm$ 0.04 & 0.11 $\pm$ 0.03\\
April 17 &38 & 0.00 & $<$ 0.06 & $<$ 0.04\\
April 18 &39 & 0.00 & 0.20 $\pm$ 0.04 & 0.10 $\pm$ 0.02\\
April 19 &40 & 0.00 & 0.24 $\pm$ 0.05 & 0.07 $\pm$ 0.03\\
April 20 &41 & 0.00 & $<$ 0.08 & $<$ 0.04\\
April 29 & 50 & 0.00 & $<$ 0.1 & $<$ 0.02 \\
May 1 & 51 & 0.00 & $<$ 0.1 & $<$ 0.02 \\
May 4 &55 & 0.00 & $<$ 0.05 & $<$ 0.02\\
May 5 &56 & 3.00 & $<$ 0.05 & $<$ 0.02\\
May 6 &57 & 0.00 & $<$ 0.05 & $<$ 0.03\\
May 11 &62 & 0.05 & $<$ 0.04 & $<$ 0.03\\
\hline
\end{tabular*}
Uncertainties from counting statistics are $\pm$ 1 standard deviation/error. Upper limits are reported at 95\% confidence limit. \\
\textsuperscript{a}The air filters which ended sampling on March 18, March 20 and March 23 were assayed in a single detector counting system and are not included here. The March 18 filter had no detectable fission product isotopes. Filters were sampled less frequently and counted for longer durations as the activity approached our detection limit.\\
\textsuperscript{b}\citep{wund}
\end{table*}
Additional fission product isotopes were detected by simultaneously counting filters from the first 10 d of air sampling at KURF. Fig.~\ref{fig:kurf} shows gamma-ray spectra from this measurement where peaks from the isotopes \textsuperscript{132}Te and \textsuperscript{136}Cs, which could not be seen in the above-ground detectors, are visible. The fission products and gamma-ray energies are listed in Table~\ref{tab:kurf}. Two individual filters were also counted at KURF to measure the activity of fission products that were below the detection limit of the above-ground detectors. The activity concentrations measured in these filters are listed in Table~\ref{tab:kurfresults}.
\begin{figure*} [htp]
\centering
\subfloat[]{\epsfig{file=Fig3a, width=.90\textwidth}}\\
\subfloat[]{\epsfig{file=Fig3b.eps, width=.45\textwidth}}
\subfloat[]{\epsfig{file=Fig3c.eps, width=.45\textwidth}}\\
\caption{(a) Gamma-ray spectrum for the 10 filters counted simultaneously (black) and background spectrum (gray) in VT-1. (b),(c) Two regions of the VT-1 filter and background spectra where various fission fragment and background peaks are visible. Number labels correspond to the isotopes listed in Table~\ref{tab:kurf}.}
\label{fig:kurf}
\end{figure*}
\begin{table*}
\small
\caption{Fission products and backgrounds detected at KURF for peaks seen in the part of the spectra displayed in Fig.~\ref{fig:kurf}.}
\label{tab:kurf}
\begin{tabular*}{.99\textwidth}{@{\extracolsep{\fill}}lllll}
\hline
\\
Isotope & Fig.~\ref{fig:kurf} Label &Source & $T_{1/2}$ & Gamma-rays detected (keV) \\
\hline
\\
\textsuperscript{132}Te & 1 & fission product & 76.9 h & 228, 326\\
\textsuperscript{214}Pb & 2 & background from \textsuperscript{222}Rn & 26.8 min & 242, 295. 352\\
\textsuperscript{131}I & 3 & fission product & 8.03 d & 80.2, 176, 284, 364, 503, 637, \\
& & & & 643, 723 \\
\textsuperscript{136}Cs & 4 & fission product & 13.0 d & 340, 818, 1048, 1235\\
\textsuperscript{134}Cs & 5 & fission product & 2.07 y & 458, 487, 563, 569, 605, 695, \\
& & & & 796, 802, 1039, 1168, 1365 \\
\textsuperscript{7}Be & 6 & cosmogenic & 53.2 d & 477\\
\textsuperscript{208}Tl & 7 & background from \textsuperscript{232}Th & 3.05 min & 583, 2614\\
\textsuperscript{214}Bi & 8 & background from \textsuperscript{222}Rn & 19.9 min & 609, 768, 1120, 1765, 2204\\
\textsuperscript{137}Cs & 9 & fission product & 30.1 y & 661\\
\textsuperscript{132}I & 10 & fission product & 137.7 min & 667, 773, 954\\
\hline
\end{tabular*}
\end{table*}
\begin{sidewaystable*}
\small
\caption{\ Assay results from filters measured at KURF. The filter efficiency for \textsuperscript{132}Te is assumed to be 99.98\%. All activity concentrations are in units of mBq/m\textsuperscript{3}.}
\label{tab:kurfresults}
\begin{tabular*}{.99\textwidth}{@{\extracolsep{\fill}}lllllll}
\hline
\\
Filter Sample & \textsuperscript{132}Te & \textsuperscript{131}I & \textsuperscript{132}I & \textsuperscript{134}Cs & \textsuperscript{136}Cs & \textsuperscript{137}Cs\\
End Date & & & & & & \\
\hline
\\
March 20 & 0.08 $\pm$ 0.03 & 0.57 $\pm$ 0.06 & 0.024 $\pm$ 0.004 & 0.038 $\pm$ 0.004 & 0.007 $\pm$ 0.002 & 0.043 $\pm$ 0.005 \\
March 29 & 0.05 $\pm$ 0.01 & 4.2 $\pm$ 0.5 & 0.08 $\pm$ 0.02 & 0.020 $\pm$ 0.003 & 0.026 $\pm$ 0.008 & 0.20 $\pm$ 0.03 \\
\\
\hline
\end{tabular*}
\end{sidewaystable*}
The maximum activity concentration detected was 4.2 $\pm$ 0.6 mBq/m\textsuperscript{3} of \textsuperscript{131}I, which did not include a correction for the volatile iodine components. If the volatile components were included, the activity concentration may have been comparable to the limit of 7.8 mBq/m\textsuperscript{3} set by the Environmental Protection Agency (EPA) although it was well below the limit of 7.4 Bq/m\textsuperscript{3} set by the Nuclear Regulatory Commission (NRC)~\citep{nrc,epa}. The NRC limit is intended to limit public dosage to less than 1 mSv effective dose.
Local weather had a significant impact on measured activities. In particular, sharp decreases in both \textsuperscript{131}I and \textsuperscript{137}Cs activity, shown in Fig.~\ref{time}, can be correlated with rain events in Table~\ref{samplingdata}.
The measurement of certain isotopes and their ratios allows us to draw some conclusions about the nature and timescale of the Fukushima accident. The measurement of short-lived isotopes, such as \textsuperscript{131,132}I, \textsuperscript{136}Cs and \textsuperscript{132}Te, indicates that radioactivity was released from recently active fuel rods and not from spent fuel cooled for long term. Isotopes such as \textsuperscript{103}Ru, \textsuperscript{141}Ce, \textsuperscript{239}Np and other fission products that were detected after the Chernobyl accident in both Europe and North America~\citep{Dev86,Per90} were not found in the data. Since almost all of the iodine entering containment from a light water reactor cooling system is in the form of CsI which is very soluble~\citep{Bea91}, this suggests that radioactivity due to the release of contaminated steam was far more abundant than from active fuel particles.
Using data from KURF, the ratio of the number of \textsuperscript{134}Cs to \textsuperscript{137}Cs atoms present in filters from the first 10 d of air sampling could also be determined. If the release of radioactivity was from a nuclear weapon, for example, there would be almost no \textsuperscript{134}Cs produced because the formation of \textsuperscript{134}Cs via neutron capture on \textsuperscript{133}Cs is strongly dependent on the length of criticality.
The atomic ratio of \textsuperscript{134}Cs to \textsuperscript{137}Cs was found to be 0.070 $\pm$ 0.001, and is consistent with a nuclear reactor accident and not a nuclear weapons test, as expected.
The ratio of \textsuperscript{131}I to \textsuperscript{137}Cs activity from a measurement of seawater samples from the South Discharge Channel, a monitoring post located 330 m south of the Discharge Channel of the unit 1 - 4 reactor sites was investigated by~\citet{Mat11}. The data were described by fitting to an exponential function. The measured ratio of of \textsuperscript{131}I to \textsuperscript{137}Cs activity from data in the current work are shown in Fig.~\ref{fig:IdivCs}. An attempt was made to fit these data to the same exponential function:
\begin{equation}
\frac{\textrm{I}_{activity}}{\textrm{Cs}_{activity}}(t) = \frac{f_\textrm{\scriptsize{I}}}{f_\textrm{\scriptsize{Cs}}}\frac{\tau_\textrm{\scriptsize{Cs}}}{\Delta t \cdot ln(2)}\left(\frac{1}{2}\right)^{t/\tau_\textrm{\scriptsize{I}}}
\end{equation}
\noindent where $f_{\textrm{\scriptsize{I}}}$ and $f_\textrm{\scriptsize{Cs}}$ are the fractions of \textsuperscript{131}I and \textsuperscript{137}Cs per fission ($2.89\times10^{-2}$ and $6.19\times10^{-2}$ respectively~\citep{ipr}), $\tau_\textrm{\scriptsize{I}}$ and $\tau_\textrm{\scriptsize{Cs}}$ are the lifetimes of \textsuperscript{131}I and \textsuperscript{137}Cs and $\Delta t$ is the length of time the reactor has been active. We assume the lifetime of \textsuperscript{131}I is short compared to the reactor operation time, the decay constant is fixed and $\Delta t$ is the only free parameter. The best fit value for $\Delta t$ was found to be 2.6 $\pm$ 0.2 y which is not in good agreement with $\Delta t$ = 1 y reported by~\citet{Mat11}. The data show large deviations from the fit ($\chi^2$/DOF = 4.4 , $p-\textrm{value}$ = $10^{-11}$) which may indicate a difference in the particulate transport mechanisms and efficiencies of the iodine and cesium isotopes.
\begin{figure}[htp]
\centering
\includegraphics[width=.75\textwidth, angle=0]{Fig4.eps}
\caption{The ratio of \textsuperscript{131}I to \textsuperscript{137}Cs activity concentration. The fit is an exponential decay where the decay constant is fixed by the lifetime of \textsuperscript{131}I.}
\label{fig:IdivCs}
\end{figure}
\section{Conclusions}
Airborne fission products released from the Fukushima Dai-ichi reactor have been measured in Chapel Hill, NC, USA. Fallout from the reactor accident is not expected to have any health implications for the people living in North Carolina or in the United States. The maximum measured \textsuperscript{131}I activity concentration is below limits set by the EPA and NRC as well as many sources of natural background radiation, including radon. After April 20, 2011, the radioactivity in the air fell below our detection limits. The measurements reported here are part of the larger global effort to quantify the transport of fallout from the Fukushima accident. We hope they will contribute to future models of the atmospheric transport of fission products.
\section{Acknowledgments}
This work was primarily supported by NSF grant \# PHY 0705014 and DOE grant numbers DE-FG02-97ER4104 and DE-FG02-97ER41033. This research was supported in part by an award from the Department of Energy (DOE) Office of Science Graduate Fellowship Program (DOE SCGF). The DOE SCGF Program was made possible in part by the American Recovery and Reinvestment Act of 2009. The DOE SCGF program is administered by the Oak Ridge Institute for Science and Education for the DOE. ORISE is managed by Oak Ridge Associated Universities (ORAU) under DOE contract number DE-AC05-06OR23100. All opinions expressed in this paper are the author's and do not necessarily reflect the policies and views of DOE, ORAU, or ORISE. We acknowledge Lhoist North America for providing us access to the underground site and logistical support. We also thank Mike Miller and Andreas Knecht for useful discussions during the preparation of this manuscript.
\clearpage
|
\section{Introduction}
Let $F$ be a finite extension of $\mathbf{Q}_p$, let $k$ be a finite field and let $V_0$ be a finite dimensional $k$-vector
space carrying a continuous representation of the absolute Galois group $G_F$. Let $\mc{O}$ be a finite totally
ramified extension of $W(k)$. There
is then a universal ring $R^{\mathrm{univ}}$ parameterizing (framed) deformations of $V_0$ to $\mc{O}$-algebras. There are
many loci in $\MaxSpec(R^{\mathrm{univ}}[1/p])$ which are of interest: for example, one has the locus of crystalline
representations with given Hodge--Tate weights.
For reasons that arose in the study of modularity lifting, one would like to be able to show that certain loci in
$\MaxSpec(R^{\mathrm{univ}}[1/p])$ are Zariski closed, and then study the corresponding reduced quotient of $R^{\mathrm{univ}}[1/p]$.
In favorable cases, one can impose evident deformation conditions and show that the resulting problem is representable
by a quotient of $R^{\mathrm{univ}}$ whose generic fiber has for its maximal ideals the locus in question; one can then study
this quotient of $R^{\mathrm{univ}}$ through its moduli-theoretic description. However, this approach is not always
viable.
Kisin was the first to obtain general results in this direction. To explain his method we must introduce some
notation. Let $X$ be a subset of $\MaxSpec(R^{\mathrm{univ}}[1/p])$. When Kisin's method is applicable, it produces a
projective morphism $\Theta : Z \to \Spec(R^{\mathrm{univ}})$ of schemes such that $\Theta[1/p]$ is a closed immersion
inducing a bijection between the closed points of $Z[1/p]$ and $X$. Furthermore, the formal completion
of $Z$ along the special fiber of $\Theta$ has an easily described functor of points. Let $\Spec(R)$ be the
scheme-theoretic image of $\Theta$. Then, with
some assumptions on $Z$ which usually hold, $R$ is the unique $\mc{O}$-flat reduced quotient of $R^{\mathrm{univ}}$ with
$\MaxSpec(R[1/p])=X$. Since $\Theta$ induces an isomorphism $Z[1/p] \to \Spec(R[1/p])$, one can study $R[1/p]$
using the moduli problem solved by the completion of $Z$.
Although the above method is well-suited to proving the existence of the ring $R$ and studying its generic fiber,
it does not yield much information about $R$ itself. For instance, $R$ is defined as a scheme-theoretic image and so
it is not at all clear how to describe its functor of points.
In this paper, we give a method for analyzing $R$ itself, in certain cases. We carry this method out in the simplest
non-trivial case,
where $V_0$ is the trivial two dimensional representation and $X$ is a certain set of ordinary representations.
Note that this is not one of the ``favorable cases'' mentioned in the second paragraph, since $V_0$ is trivial.
We are able to show that $R$ (or closely related rings) is normal, Cohen--Macaulay and not Gorenstein. The
Cohen--Macaulay property has well-known
global consequences: it implies that global deformation rings using this local condition are torsion-free, and can be
used to improve Kisin's $R[1/p]=\mathbf{T}[1/p]$ theorem to an $R=\mathbf{T}$ theorem. We believe that the method we use here
should work for other two dimensional ordinary deformations, and may be feasible for higher dimensional ordinary
representations. It is not clear to us if it will be useful in other cases, however.
\subsection{Outline of the method in general}
\label{ss:out}
Let $X$ be an equidimensional Zariski closed subset of $\MaxSpec(R^{\mathrm{univ}}[1/p])$ and let $R$ be the unique
$\mc{O}$-flat reduced quotient of $R^{\mathrm{univ}}$ such that $\MaxSpec(R[1/p])=X$. To analyze $R$ we proceed along
the following steps:
\begin{enumerate}
\item Come up with a list of equations which make sense in arbitrary deformations of $V_0$ and which hold on the set
$X$. Let $R^{\dag}$ be the quotient of $R^{\mathrm{univ}}$ by these equations.
\item Show that the map $R^{\mathrm{univ}} \to R$ factors through $R^{\dag}$, and that the resulting map $R^{\dag}[1/p]_{\mathrm{red}}
\to R[1/p]$ is an isomorphism. This should not be difficult if one found enough equations in step (a).
One should understand $R[1/p]$ from Kisin's method, and so one should now have some understanding of $R^{\dag}[1/p]$.
\item Show that $R^{\dag} \otimes_{\mc{O}} k$ is reduced, equidimensional of the same dimension as $R^{\dag}[1/p]$ and
has the same number of minimal primes as $R^{\dag}[1/p]$. As far as we know, there is no reason that this must be
true; however, if it is true, it may be tractable to prove, as $R^{\dag} \otimes_{\mc{O}} k$ represents an explicit
deformation problem on $k$-algebras.
\item Appeal to abstract commutative algebra facts to conclude that $R^{\dag}$ is $\mc{O}$-flat and reduced
(see Propositions~\ref{flat} and~\ref{reduced}).
\item Conclude that $R^{\dag} \to R$ is an isomorphism, as both are $\mc{O}$-flat and reduced and the map
$R^{\dag}[1/p]_{\mathrm{red}} \to R[1/p]$ is an isomorphism.
\end{enumerate}
If the above steps can be carried out then one has a moduli-theoretic description of $R$, namely, the equations used
to define $R^{\dag}$. This can then be used to study $R$.
\subsection{Outline of the method in a specific case}
\label{ss:out2}
Let us give a more detailed outline of the above method in a specific, somewhat easy, case. Let $V_0$ be the trivial
two dimensional representation and let $X$ be the set of representations which are conjugate to one of the form
\begin{displaymath}
\mat{\chi}{\ast}{}{1},
\end{displaymath}
where $\chi$ denotes the cyclotomic character. We assume that $\chi=1 \pmod{p}$, otherwise $X$ is empty. There
are two obvious families of equations that make sense in any deformation and that hold on $X$, namely:
\begin{itemize}
\item $\tr(g)=\chi(g)+1$ for $g \in G_F$.
\item $(g-1)(g'-1)=(\chi(g)-1)(g'-1)$ for $g, g' \in G_F$.
\end{itemize}
Let $R^{\dag}$ be the quotient of $R^{\mathrm{univ}}$ by these equations. To be more precise, let $\rho^{\mathrm{univ}}:G_F \to
\mathrm{GL}_2(R^{\mathrm{univ}})$ be the universal framed deformation. Let $I$ be the ideal of $R^{\mathrm{univ}}$ generated by
the elements
\begin{displaymath}
\tr(\rho^{\mathrm{univ}}(g))-\chi(g)-1,
\end{displaymath}
as $g$ varies over $G_F$, as well as the entries of the matrix
\begin{displaymath}
(\rho^{\mathrm{univ}}(g)-1)(\rho^{\mathrm{univ}}(g')-1)-(\chi(g)-1)(\rho^{\mathrm{univ}}(g')-1),
\end{displaymath}
as $g$ and $g'$ vary over $G_F$. Then $R^{\dag}=R^{\mathrm{univ}}/I$. Thus $R^{\dag}$ is described as a quotient of
$R^{\mathrm{univ}}$ by an explicit, though uncountable, set of relations. Nonetheless, it is clear what the functor of
points of $R^{\dag}$ is.
The only really interesting remaining step in the above procedure is (c), i.e., the analysis of the ring
$R^{\dag} \otimes_{\mc{O}} k$. To understand this ring, it suffices to understand maps $R^{\dag} \to A$ where
$A$ is a $k$-algebra. From the equations defining $R^{\dag}$, together with the fact that $\chi=1\pmod{p}$, we
see that a map $R^{\mathrm{univ}} \to A$, corresponding to a deformation $V$, factors through $R^{\dag}$ if and only if
$\tr(g \vert_V)=2$ for all $g \in G_F$ and $(g-1)(g'-1)V=0$ for all $g, g' \in G_F$. It is easy to see that
the action of $G_F$ on such a deformation $V$ factors through the abelianization of $G_F$.
Using class field theory, it is therefore possible to give a very explicit description of $\Spec(R^{\dag}
\otimes_{\mc{O}} k)$ as a certain space of tuples of matrices.
(In fact, it is (a formal completion of) the space $\mc{A}_{d+2}$ discussed in \S \ref{aspace}, with $d=[F:\mathbf{Q}_p]$.)
This space can be analyzed by techniques of algebraic geometry and seen to be integral, normal and Cohen--Macaulay.
All the fine points in the above method carry through in this case, and so the map $R^{\dag} \to R$ is an isomorphism.
This gives a moduli-theoretic description of $R$: a map $R \to A$ corresponds to a framed deformation of $V_0$ to $A$
where the obvious equations given above hold. Furthermore, since we know that $R^{\dag}$ is $\mc{O}$-flat and
$R^{\dag} \otimes_{\mc{O}} k$ is Cohen--Macaulay, we find that $R$ itself is Cohen--Macaulay. A similar argument can
be used to show that $R$ is normal.
\subsection{Generalizations}
There are two natural directions in which one could attempt to extend the results of this paper. First, one could
remain in the two dimensional case but work with non-ordinary representations. This seems like it may be
difficult, as it is probably not easy to determine the equations that hold on $X$.
Second, one could remain in the ordinary case but work with higher dimensional representations.
This seems to be a more tractable avenue, since it is easy to come up with many equations that hold on $X$.
Analyzing the special fiber of $R^{\dag}$ may be difficult, however.
\subsection{Plan of paper}
We begin in \S \ref{s:alg} by establishing some miscellaneous results we will need.
In \S \ref{s:modmat}, we study certain moduli spaces of matrices. These spaces will show up as the
special fibers of the rings $R^{\dag}$ (as we saw in \S \ref{ss:out2}), and it is crucial to know that they are
reduced and have the appropriate number of irreducible components. In
\S \ref{s:local} we carry out the analysis of ordinary deformation rings in the local case and prove the main results
of the paper. Finally, in \S \ref{s:global}, we give the applications to global deformation rings.
\subsection*{Acknowledgements}
I would like to thank Bhargav Bhatt, Brian Conrad and Mark Kisin for useful conversations. I would also like to
thank Steven Sam for his comments on a draft of this paper.
\section{Preliminary results}
\label{s:alg}
In this section we establish some preliminary results that we will need later on. These are all likely well-known
and quite possibly in the literature already. However, we do not know convenient references for many of them, and
so include proofs.
\subsection{Rings associated with vector bundles}
Let $X$ be a geometrically integral scheme proper over $k$, let $\eta$ be a vector bundle on $X$ and let $Z$ be the
total space of the dual of $\eta$. Put
\begin{displaymath}
R=\Gamma(Z, \mc{O}_Z)=\Gamma(X, \Sym(\eta)).
\end{displaymath}
In this section we prove two results about $R$, the first concerning its singularities and the second its minimal
free resolution. We will only apply these results when $X=\mathbf{P}^1$, but the proofs are just as easy in the general
case. The first result is the following (the notation is explained following the proposition):
\begin{proposition}
\label{geo-1}
Assume that $X$ is normal and Cohen--Macaulay, that $\eta$ is ample and generated by its global sections and that
\begin{displaymath}
H^i(X, \Sym(\eta))=H^i(X, \Sym(\eta) \otimes \det(\eta) \otimes \omega_X)=0
\end{displaymath}
for $i>0$. Then the map $Z \setminus Z_0 \to \Spec(R) \setminus \{0\}$ is an isomorphism and $R$ is Cohen--Macaulay.
Furthermore, if
\begin{displaymath}
\dim{H^0(X, \det(\eta) \otimes \omega_X)}>1
\end{displaymath}
then the local ring of $R$ at 0 is not Gorenstein.
\end{proposition}
Let us explain the notation and terminology used above. The ring $R$ is naturally graded. Let $R_+$ be the ideal of
positive degree elements. Then $R/R_+$ is identified with $k$. We write 0 for the point in $\Spec(R)$ corresponding
to the ideal $R_+$. We write $\omega_Y$ for the dualizing complex of a variety $Y$ over $k$, which is a coherent
sheaf when $Y$ is Cohen--Macaulay.
A vector bundle $\eta$ on $X$ is said to be \emph{ample} if the line bundle $\mc{O}(1)$ on $\mathbf{P}(\eta^{\vee})$ is ample.
If $\eta$ is a line bundle then $\mathbf{P}(\eta^{\vee})=X$ and $\mc{O}(1)$ is just $\eta$, and so this notion of ample
corresponds to the usual one. A direct sum of ample bundles is again ample. We write $Z_0$ for
the image of the zero section of $\pi:Z \to X$.
Before proving the proposition we give some lemmas. In these lemmas, we do not assume the hypotheses of the
proposition.
\begin{lemma}
\label{geo-1a}
If $\eta$ is generated by global sections then $Z \to \Spec(R)$ is proper.
\end{lemma}
\begin{proof}
Since $\eta$ is generated by its global sections, it is a quotient of $V^* \otimes \mc{O}_X$ for some vector
space $V$. Thus $Z$ is a closed subset of $X \times V$. As the map $X \times V \to V$ is proper, we find that
the map $Z \to V$ is proper. This map factors as $Z \to \Spec(R) \to V$, and so $Z \to \Spec(R)$ is proper.
\end{proof}
\begin{lemma}
\label{geo-1b}
Assume that $X$ is normal, that $\eta$ is ample and that $f:Z \to \Spec(R)$ is proper. Then $f$ induces an
isomorphism $Z \setminus Z_0 \to (\Spec{R}) \setminus \{0\}$.
\end{lemma}
\begin{proof}
Since $\eta$ is ample,
it follows that, for $z \in Z(\ol{k})$, we have $f(z)=0$ if and only if $z \in Z_0(\ol{k})$. Furthermore,
if $z \not\in Z_0(\ol{k})$ then one can recover $\pi(z)$ from $f(z)$; since the restriction of $f$ to each fiber of
$\pi$ is injective on $\ol{k}$-points, it follows that the restriction of $f$ to
$Z \setminus Z_0$ is injective on $\ol{k}$-points.
Let $f'$ denote the map $Z \setminus Z_0 \to (\Spec{R}) \setminus \{0\}$ induced by $f$. As $f$ is proper, so too is
$f'$. We thus find that the image of $f'$ is a closed subscheme of $(\Spec{R}) \setminus \{0\}$. Since $R$ is
integral and has dimension at most that of $Z$, and $f'$ is injective on $\ol{k}$-points, it follows that $f'$
is surjective. It follows that $f'$ is finite (Zariski's main theorem) and birational (as it is
generically flat), which implies (by normality) that $f'$ is an isomorphism.
\end{proof}
\begin{lemma}
\label{geo-1c}
Let $f:\wt{Y} \to Y$ be a proper birational map of schemes over $k$, with $\wt{Y}$ Cohen--Macaulay and
$Y$ affine. Suppose that $H^i(\wt{Y}, \mc{O}_{\wt{Y}})=H^i(\wt{Y}, \omega_{\wt{Y}})=0$
for $i>0$ and that the natural map $f^*:H^0(Y, \mc{O}_Y) \to H^0(\wt{Y}, \mc{O}_{\wt{Y}})$ is an isomorphism. Then $Y$
is Cohen--Macaulay and $\omega_Y$ is isomorphic to $f_*(\omega_{\wt{Y}})$.
\end{lemma}
\begin{proof}
Grothendieck duality for $f$ states that
\begin{displaymath}
Rf_* R\uHom_{\wt{Y}}(F, f^!G)=R\uHom_Y(Rf_* F, G)
\end{displaymath}
for $F \in D^b(\wt{Y})$ and $G \in D^b(Y)$. Take $F=\mc{O}_{\wt{Y}}$ and $G=\omega_Y$. The hypotheses of the lemma
imply that $Rf_* F=\mc{O}_Y$; furthermore, $f^!G=\omega_{\wt{Y}}$. We thus find $Rf_*(\omega_{\wt{Y}})=\omega_Y$. The
hypotheses of the lemma imply that $Rf_*(\omega_{\wt{Y}})=f_*(\omega_{\wt{Y}})$. Thus $\omega_Y$ is concentrated in
a single degree, and so $Y$ is Cohen--Macaulay.
\end{proof}
We now return to the proof of Proposition~\ref{geo-1}.
\begin{proof}[Proof of Proposition~\ref{geo-1}]
The scheme $Z$ is geometrically integral and Cohen--Macaulay, being a vector bundle over such a scheme. By
hypothesis, $H^i(Z, \mc{O}_Z)=H^i(X, \Sym(\eta))=0$ for $i>0$. We have
$\omega_Z=\pi^*(\det(\eta) \otimes \omega_X)$; when $X$ is smooth, this can be seen from taking the determinant of
the exact sequence
\begin{displaymath}
0 \to \pi^*(\Omega^1_{X/k}) \to \Omega^1_{Z/k} \to \Omega^1_{Z/X} \to 0
\end{displaymath}
after using the identification $\Omega^1_{Z/X}=\pi^*(\eta)$. Thus
\begin{displaymath}
H^i(Z, \omega_Z)=H^i(X, \Sym(\eta) \otimes \det(\eta) \otimes \omega_X)
\end{displaymath}
vanishes for $i>0$, by hypothesis. From Lemmas~\ref{geo-1a} and~\ref{geo-1b}, we see that
$Z \to \Spec(R)$ is proper and birational. Lemma~\ref{geo-1c} shows that $R$ is Cohen--Macaulay and
$\omega_R=H^0(Z, \omega_Z)$. Now, $\omega_R$ is a graded $R$-module whose first graded piece is
$H^0(X, \det(\eta) \otimes \omega_X)$. This space injects into $\omega_R/R_+ \omega_R$. Thus if $H^0(X,
\det(\eta) \otimes \omega_X)$ has dimension greater than 1 then the fiber of $\omega_R$ at 0 has dimension greater
than 1 and $R$ is not Gorenstein at 0.
\end{proof}
We now turn towards our second result. Suppose that $\eta$ is generated by its global sections, and write
\begin{displaymath}
0 \to \xi \to \epsilon \to \eta \to 0
\end{displaymath}
with $\epsilon$ a globally free coherent $\mc{O}_X$-module. Let
$V=\Hom(\epsilon, \mc{O}_X)$, so that we have canonical identifications $\epsilon=V^* \otimes \mc{O}_X$ and
$\Spec(\Sym(\epsilon))=X \times V$. Put $S=\Sym(V^*)$. Note that there is a natural map $S \to R$ respecting the
grading. We regard $k$ as an $S$-module via the identification $S/S_+=k$.
Our second result is the following proposition. It is the basis of Weyman's ``geometric method'' for studying
syzygies, as exposited in \cite[Ch.~5]{Weyman}. We include a proof since it is short.
\begin{proposition}
\label{geo-2}
Assume $H^i(X, \Sym(\eta))=0$ for $i>0$. Then we have an isomorphism of graded vector spaces
\begin{displaymath}
\Tor^n_S(R, k)=\bigoplus_{i \ge n} H^{i-n}(X, \lw{i}{\xi})[i],
\end{displaymath}
where $[ \cdot ]$ indicates the grading.
\end{proposition}
\begin{proof}
Consider the diagram
\begin{displaymath}
\xymatrix{
X \ar[r]^-{i'} \ar[d]_{p'} & X \times V \ar[d]^p \\
\ast \ar[r]^i & V }
\end{displaymath}
where $\ast=\Spec(k)$ and the horizontal maps are zero sections. We have isomorphisms (in the derived category of
graded $k$-vector spaces)
\begin{displaymath}
\Tor^{\bullet}_S(R, k)=Li^* Rp_* \mc{O}_Z=Rp'_* L(i')^* \mc{O}_Z.
\end{displaymath}
The first isomorphism comes from the fact that $Rp_* \mc{O}_Z=p_* \mc{O}_Z=R$, while the second
is the base change map, which is an isomorphism in this case (as can be seen by applying the projection formula of
\cite[Prop.~5.6]{Hartshorne}, taking $F=\mc{O}_Z$ and $G$ the structure sheaf of the point).
Now, the Koszul complex gives a resolution of $\Sym(\eta)$ as a $\Sym(\epsilon)$-module:
\begin{displaymath}
\cdots \to \Sym(\epsilon) \otimes \lw{2}{\xi} \to \Sym(\epsilon) \otimes \lw{1}{\xi} \to \Sym(\epsilon) \to
\Sym(\eta) \to 0.
\end{displaymath}
The terms of this resolution are graded $\Sym(\epsilon)$-modules, with $\lw{i}{\xi}$ being in degree $i$. The
differentials have degree one. Let $q:X \times V \to X$ be the projection. We can recast the above
resolution as a quasi-isomorphism of complexes of coherent $\mc{O}_{X \times V}$ modules:
\begin{displaymath}
[ \lw{\bullet}(q^*\xi) ] \to \mc{O}_Z,
\end{displaymath}
As the sheaves in the Koszul complex are locally free $\mc{O}_{X \times V}$-modules, we can calculate $L(i')^*$ by
simply applying $(i')^*$. After doing so all differentials vanish, since the differentials in the Koszul complex
have degree one. We thus have a quasi-isomorphism
\begin{displaymath}
[ \lw{\bullet}(\xi) ] \to L(i')^* \mc{O}_Z
\end{displaymath}
where the complex on the left has zero differentials. Applying $Rp'_*$ yields the formula in the statement of the
proposition.
\end{proof}
\subsection{A flatness criterion}
The main result of this section gives a fiberwise criterion for flatness over a discrete valuation ring. Before
stating it, let us recall a few definitions. Let $A$ be a noetherian ring. We say that $A$ is \emph{catenary} if for
any two
prime ideals $\mf{p} \subset \mf{q}$ of $A$, any two maximal chains of prime ideals beginning at $\mf{p}$ and ending at
$\mf{q}$ have the same length. This is a mild condition satisfied by most rings one encounters; for instance, every
finitely generated algebra over a complete local noetherian ring is catenary. The \emph{dimension}
(resp.\ \emph{codimension}) of a prime ideal $\mf{p}$ of $A$ is
the length of the longest chain of primes beginning (resp.\ ending) at $\mf{p}$, or, equivalently, the dimension
of $A/\mf{p}$ (resp.\ $A_{\mf{p}}$). We say that $A$ is \emph{equidimensional} of dimension $d$ if $\dim(\mf{p})
+\codim(\mf{p})=d$ for all prime ideals $\mf{p}$. If $A$ is a noetherian catenary local ring then $A$ is
equidimensional if and only if its minimal primes all have the same dimension. We can now state the proposition:
\begin{proposition}
\label{flat}
Let $A$ be a catenary noetherian local ring and let $\pi$ be an element of the maximal ideal of $A$. Assume the
following:
\begin{itemize}
\item The ring $A/\pi A$ is reduced.
\item The rings $A[1/\pi]$ and $A/\pi A$ are equidimensional of the same dimension and have the same number of
minimal primes.
\end{itemize}
Then $\pi$ is not a zero-divisor in $A$.
\end{proposition}
\begin{remark}
Let $\mc{O}$ be a discrete valuation ring with uniformizer $\pi$ and let $A$ be a catenary noetherian local
$\mc{O}$-algebra. The above proposition gives a criterion for $A$ to be flat over $\mc{O}$ in terms of conditions on
the fibers of the map $\Spec(A) \to \Spec(\mc{O})$.
\end{remark}
\begin{remark}
The proof of the proposition will yield the following additional piece of information: extension gives a bijection
between the minimal primes of $A$ and those of $A[1/\pi]$, and similarly for $A/\pi A$ in place of $A[1/\pi]$.
\end{remark}
We now prove the proposition. If $\pi$ is nilpotent then the proposition is trivial, so assume this is not
the case. Let $d$ be the common dimension of $A[1/\pi]$ and $A/\pi A$, and let $n$ be the common number of minimal
primes in $A[1/\pi]$ and $A/\pi A$. Let $I$ be the ideal of $\pi$-power torsion in $A$. We must show that $I=0$. Put
$B=A/I$. Note that $B$ is still catenary, noetherian and local, and $\pi$ is not a zero-divisor in $B$.
\begin{lemma}
\label{flat-1}
No minimal prime of $B$ contains $\pi$.
\end{lemma}
\begin{proof}
Let $\mf{p}_1, \ldots, \mf{p}_n$ be the minimal primes of $B$, and suppose $\pi$ belongs to $\mf{p}_1$. We
cannot have $n=1$, as $\pi$ is not nilpotent. Let $x_i$, for $2 \le i \le n$, be an element of $\mf{p}_i$ which
does not belong to $\mf{p}_1$, and let $x$ be the product of the $x_i$. Thus $x$ belongs to $\mf{p}_2 \cap \cdots
\cap \mf{p}_n$, but is not nilpotent. As $\pi x$ is nilpotent, we find that $\pi^k$ kills $x^k$ for some $k$, which
contradicts the fact that $\pi$ is not a zero-divisor of $B$.
\end{proof}
\begin{lemma}
\label{flat-2}
Any prime of $B$ minimal over $(\pi)$ has codimension 1.
\end{lemma}
\begin{proof}
Krull's principal ideal theorem says the codimension is at most 1, while Lemma~\ref{flat-1} says it cannot be 0.
\end{proof}
\begin{lemma}
\label{flat-3}
Let $R$ be a noetherian local domain and let $\pi$ be a non-zero element of the maximal ideal of $R$ such that
$R[1/\pi]$ is a field. Then $R$ has exactly two prime ideals: its maximal ideal and the zero ideal.
\end{lemma}
\begin{proof}
Since $R[1/\pi]$ is a field, given any non-zero $x \in R$, we can find a non-zero $y \in R$ such that $xy=\pi^n$ for
some $n$. It follows that every non-zero principal ideal, and therefore every non-zero ideal, contains a power of
$\pi$. Thus $\pi$ belongs to every non-zero prime ideal. We thus see that the codimension one primes of $R$ are in
bijection with the minimal primes of $R/\pi R$, and are thus finite in number.
Let $\mf{p}_1, \ldots, \mf{p}_n$ be the codimension one primes of $R$. Every element of the maximal ideal $\mf{m}$ of
$R$ belongs to one of these primes, by Krull's principal ideal theorem. Thus $\mf{m} \subset \bigcup \mf{p}_i$.
However, this implies $\mf{m} \subset \mf{p}_i$ for some $i$, and so there is only one $\mf{p}_i$ and it is maximal.
This completes the proof.
\end{proof}
\begin{lemma}
\label{flat-4}
The ring $B$ is equidimensional of dimension $d+1$ and contains $n$ minimal primes.
\end{lemma}
\begin{proof}
The primes of $B$ not containing $\pi$ correspond to the primes of $B[1/\pi]=A[1/\pi]$ via extension and contraction.
Since no minimal prime of $B$ contains $\pi$, it follows that the minimal primes of $B$ and $B[1/\pi]$ are in bijection,
and so $B$ has $n$ minimal primes. Let $\mf{p}_0$ be a minimal prime of $B$, and let $\mf{q}_0$ be its extension to
$B[1/\pi]$. Let $\mf{q}_0 \subset \cdots \subset \mf{q}_d$ be a maximal chain of prime ideals in $B[1/\pi]$ and
let $\mf{p}_i$ be the contraction of $\mf{q}_i$. Then $B/\mf{p}_d$ is a local domain in which $\pi$ is non-zero
element of the maximal ideal, and $(B/\mf{p}_d)[1/\pi]=B[1/\pi]/\mf{q}_d$ is a field. It follows from
Lemma~\ref{flat-3} that $B/\mf{p}_d$ is one dimensional, and so there is no prime between $\mf{p}_d$ and the maximal
ideal $\mf{p}_{d+1}$ of $B$. Thus
$\mf{p}_0 \subset \cdots \subset \mf{p}_{d+1}$ is a maximal chain of length $d+1$. As $B$ is catenary, every
maximal chain between $\mf{p}_0$ and $\mf{p}_{d+1}$ has length $d+1$, and so $B$ is equidimensional of dimension $d+1$.
\end{proof}
\begin{lemma}
\label{flat-5}
The ring $B/\pi B$ is equidimensional of dimension $d$.
\end{lemma}
\begin{proof}
Let $\mf{p}$ be a minimal prime of $B/\pi B$ and let $\wt{\mf{p}}$ be its inverse image in $B$. Then $\wt{\mf{p}}$
has codimension 1 by Lemma~\ref{flat-2}, and since $B$ is equidimensional, dimension $d$. Of
course, $\wt{\mf{p}}$ and $\mf{p}$ have the same dimension. This shows that all minimal primes of $B/\pi B$ have
dimension $d$, and as $B/\pi B$ is catenary, noetherian and local, the proposition follows.
\end{proof}
\begin{lemma}
\label{flat-6}
Let $\mf{p}$ be a minimal prime of $B$. Then there exists a prime $\mf{q}$ of $B$ minimal over $(\pi)$ which
contains $\mf{p}$.
\end{lemma}
\begin{proof}
The ring $B/\mf{p}$ is a local domain of dimension $d+1$ in which $\pi$ is non-zero element of the maximal ideal, and
so $B/((\pi)+\mf{p})$ has dimension $d$. Let $\mf{q}$ be the inverse image in $B$ of a minimal prime of
$B/((\pi)+\mf{p})$ of dimension $d$. Then $\mf{q}$ has dimension $d$, and since $B$ is equidimensional, codimension 1.
It follows from Lemma~\ref{flat-2} that $\mf{q}$ is minimal over $(\pi)$.
\end{proof}
\begin{lemma}
\label{flat-7}
The ring $B/\pi B$ is regular in codimension 0, that is, if $\mf{q}$ is a minimal prime then $(B/\pi B)_{\mf{q}}$ is
a field.
\end{lemma}
\begin{proof}
Let $\mf{q}$ be a minimal prime of $B/\pi B$ and let $\mf{p}$ be its inverse image in $A/\pi A$. Since $A/\pi A$ and
$B/\pi B$ are equidimensional of the same dimension, $\mf{p}$ is a minimal prime of $A/\pi A$. One readily verifies
that the natural map $(A/\pi A)_{\mf{p}} \to (B/\pi B)_{\mf{q}}$ is a surjection of rings. As $A/\pi A$
is reduced, the former is a field, and so the latter is as well.
\end{proof}
\begin{lemma}
\label{flat-8}
The ring $B/\pi B$ contains at least $n$ minimal primes.
\end{lemma}
\begin{proof}
Let $\mf{q}$ be a minimal prime over $(\pi)$ in $B$. We claim that $\mf{q}$ contains exactly one minimal prime of $B$.
Since $(B/\pi B)_{\mf{q}}=B_{\mf{q}}/\pi B_{\mf{q}}$ is a field (Lemma~\ref{flat-7}), we find that $\pi B_{\mf{q}}$ is
the maximal ideal of $B_{\mf{q}}$. Since the maximal ideal of $B_{\mf{q}}$ is principal, any other prime ideal of
$B_{\mf{q}}$ is the zero ideal, which establishes the claim. Since every minimal prime of $B$ is contained in at least
one prime minimal over $(\pi)$ (Lemma~\ref{flat-6}), it follows that there are at least $n$ primes of $B$ minimal over
$(\pi)$.
\end{proof}
\begin{lemma}
\label{flat-9}
The map $A/\pi A \to B/\pi B$ is an isomorphism.
\end{lemma}
\begin{proof}
It is a map of equidimensional noetherian rings of the same dimension such that the source is reduced and the target
has at least the number of minimal primes as the source.
\end{proof}
\begin{lemma}
\label{flat-10}
We have $I=0$.
\end{lemma}
\begin{proof}
Since $B$ has no $\pi$-torsion, the kernel of the map $A/\pi A \to B/\pi B$ is $I/\pi I$. Thus $I/\pi I=0$, and so
$I=0$ by Nakayama's lemma.
\end{proof}
\subsection{Two more results from commutative algebra}
Let $A$ be a ring and let $\pi$ be an element of $A$. In this section we give a pair of results which allow us to
transfer properties of $A/\pi A$ and $A[1/\pi]$ to $A$.
\begin{proposition}
\label{reduced}
Suppose that $\pi$ is not a zero-divisor, $A$ is $(\pi)$-adically separated and $A/\pi A$ is reduced. Then $A$ is
reduced.
\end{proposition}
\begin{proof}
Suppose $x_1$ is an element of $A$ such that $x_1^n=0$. Then $x_1=0$ in $A/\pi A$, and so we can write $x_1=\pi x_2$.
We thus find $\pi^n x_2^n=0$, and so, since $\pi$ is not a zero-divisor, $x_2^n=0$. Reasoning as before, we can
write $x_2=\pi x_3$. Continuing in this manner, we find that $x_1$ belongs to $(\pi^n)$ for all $n$, and is thus equal
to 0. This completes the proof.
\end{proof}
\begin{proposition}
\label{normal}
Suppose that $A$ is a domain, $A[1/\pi]$ is normal and $A/\pi A$ is reduced. Then $A$ is normal.
\end{proposition}
\begin{proof}
Let $x$ be an element of the fraction field of $A$ which is integral over $A$; we must show that $x$ belongs to $A$.
Since $A[1/\pi]$ is normal, $x$ belongs to $A[1/\pi]$, and so we can write $x=y/\pi^n$ with $y \in A$ and $n \ge 0$
minimal. Assume $n>0$. Let $\sum_{i=0}^d a_i x^i$ be a monic equation over $A$ satisfied
by $x$. We then have $\sum_{i=0}^d a_i y^i (\pi^n)^{d-i}=0$. We find $y^d=0$ in $A/\pi A$, which shows that $y$
belongs to $(\pi)$, contradicting the minimality of $n$. Thus $n=0$ and $x$ belongs to $A$.
\end{proof}
\subsection{A result from Galois theory}
Let $F/\mathbf{Q}_p$ be a finite extension of degree $d$ and let $F'/F$ be the maximal $p$-power Galois extension in which the
inertia group is abelian and killed by $p$. Let $G=\Gal(F'/F)$ and let $U$ be the inertia subgroup of $G$. We thus
have a short exact sequence
\begin{displaymath}
1 \to U \to G \to \mathbf{Z}_p \to 0,
\end{displaymath}
where $\mathbf{Z}_p$ has for a topological generator the arithmetic Frobenius element $\phi$. We give $U$ the structure of
an $\mathbf{F}_p \lbb T \rbb$-module by letting $T$ act by $\phi-1$. The following result determines the structure of $U$:
\begin{proposition}
\label{galstruct}
If $F$ contains the $p$th roots of unity then
$U \cong \mathbf{F}_p \oplus \mathbf{F}_p \lbb T \rbb^{\oplus d}$; otherwise $U \cong \mathbf{F}_p \lbb T \rbb^{\oplus d}$
\end{proposition}
\begin{proof}
Fix an element $\wt{\phi}$ in $G$ lifting $\phi$.
Let $F_n$ be the unramified extension of $F$ of degree $p^n$, and let $F_n'$ be the maximal abelian extension of $F_n$
of exponent $p$ on which $\wt{\phi}^{p^n}$ acts trivially. Then $F_n'$ is the fixed field of the normal closure of
$\langle \wt{\phi} \rangle \subset G$ acting on $F'$, and so $\Gal(F_n'/F_n)$ is identified with $U/T^{p^n} U$. By
class field theory,
the abelianized Galois group of $F_n$ is identified with $(F_n^{\times})^{\wedge}$, with the element $\wt{\phi}^{p^n}$
corresponding to some element $x_n$ of valuation 1. It follows that we have isomorphism
\begin{displaymath}
\Gal(F_n'/F_n)=(F_n^{\times})^{\wedge}/\langle x_n \rangle \otimes \mathbf{F}_p=U_{F_n} \otimes \mathbf{F}_p,
\end{displaymath}
where $U_{F_n}$ is the unit group of $F_n$. We thus find that $U/T^{p^n} U$ has dimension $\epsilon+p^n d$ over
$\mathbf{F}_p$, where $\epsilon$ is 1 or 0 according to whether $F$ contains the $p$th roots of unity or not. Now, since $U$
is the
inverse limit of the $U/T^{p^n} U$ and $U/TU$ is finite dimensional, it follows that $U$ is finitely generated over
$\mathbf{F}_p \lbb T \rbb$. Appealing to the structure theory of finitely generated $\mathbf{F}_p \lbb T \rbb$-modules and our
formula for the dimension of $U/T^{p^n} U$ gives the stated result.
\end{proof}
\begin{remark}
Suppose that $F$ contains the $p$th roots of unity. By the above result, $U$ contains a unique $\mathbf{F}_p$-line on which
$\phi$ acts trivially. Let us now describe this line more explicitly. As $F_n$ contains the $p$th roots of unity,
there is some element $y_n$ of $F_n^{\times} \otimes \mathbf{F}_p$ such that $F_{n+1}=F_n(y_n^{1/p})$. This element is
unique up to scaling by elements of $\mathbf{F}_p^{\times}$. Alternatively, fixing a $p$th root of unity and letting
$(,)$ be the $\mathbf{F}_p$-valued Hilbert symbol on $F_n^{\times} \otimes \mathbf{F}_p$, we can characterize $y_n$ uniquely by
$(y_n, x)=\val(x)$. The element $y_n$ is a unit, invariant under $\phi$ and satisfies $N(y_{n+1})=y_n$, where
$N:F_{n+1}^{\times} \to F_n^{\times}$ is the norm map. The sequence
$(y_n)$ defines an element of the inverse limit of the $U_{F_n} \otimes \mathbf{F}_p$ (where the transition maps are the
norm maps). By the above proof, this inverse limit is $U$. Since $(y_n)$ is non-zero and $\phi$-invariant, it spans
the unique $\phi$-invariant $\mathbf{F}_p$-line in $U$.
\end{remark}
\section{Moduli spaces of matrices}
\label{s:modmat}
In this section we study three moduli problems related to $2 \times 2$ nilpotent matrices. Certain local rings of
these spaces will later be identified as the special fiber of certain Galois deformation rings. Throughout, $k$ is a
fixed field of characteristic not 2.
\subsection{Borel and nilpotent matrix algebras}
Let $\mf{g}=M_2$ be the space of $2 \times 2$ matrices over $k$ and let $\mf{g}^{\circ}$ be the subspace of traceless
matrices. Since 2 is invertible, the space $\mf{g}$ is the direct sum of $\mf{g}^{\circ}$ and the space of scalar
matrices. We write $\ul{\mf{g}}$ and $\ul{\mf{g}}^{\circ}$ for the sheaves $\mf{g} \otimes \mc{O}$ and $\mf{g}^{\circ}
\otimes \mc{O}$ on $\mathbf{P}^1$.
Let $T$ be a $k$-scheme. A \emph{nilpotent subalgebra} of $\mf{g}_T$ is a line subbundle $\mf{u}$ of $\mf{g}_T$
such that $\ker(\mf{u})=\im(\mf{u})$ is a line subbundle of $\mc{O}_T^2$. One easily sees that sending $\mf{u}$ to
$\ker(\mf{u})$ defines a bijection between nilpotent subalgebras of $\mf{g}_T$ and line subbundles of $\mc{O}_T^2$.
One recovers $\mf{u}$ from a line bundle $\mc{L}$ via the formula $\mf{u}=\uHom(\mc{O}_T^2/\mf{L}, \mc{L})$. Since
the bundle $\mc{O}(-1)$ on $\mathbf{P}^1$ is the universal line subbundle of $\mc{O}^2$, it follows that the nilpotent
subalgebra $\ul{\mf{u}}=\uHom(\mc{O}^2/\mc{O}(-1), \mc{O}(-1)) \cong \mc{O}(-2)$ of $\ul{\mf{g}}$ on $\mathbf{P}^1$ is the
universal nilpotent subalgebra of $\mf{g}$.
A \emph{Borel subalgebra} of $\mf{g}_T$ is a rank three subbundle $\mf{b}$ of $\mf{g}_T$ for which there
exists a line subbundle $\mc{L}$ of $\mc{O}_T^2$ such that $\mf{b} \mc{L} \subset \mc{L}$. Again, one finds that
the correspondence between $\mf{b}$ and $\mc{L}$ is bijective, and so there is a universal Borel subalgebra
$\ul{\mf{b}}$ of $\ul{\mf{g}}$ on $\mathbf{P}^1$. One can make similar definitions in the traceless case, and obtain a
universal
Borel subalgebra $\ul{\mf{b}}^{\circ}$ of $\ul{\mf{g}}^{\circ}$ on $\mathbf{P}^1$. We have $\ul{\mf{b}}=\ul{\mf{b}}^{\circ}
\oplus \mc{O}$ since 2 is invertible. As every section of $\ul{\mf{b}}^{\circ}$ induces an endomorphism of
$\mc{O}^2/\mc{O}(-1)$, we get a canonical map $\ul{\mf{b}}^{\circ} \to \mc{O}$, the kernel of which is $\ul{\mf{u}}$.
That is, we have an exact sequence
\begin{displaymath}
0 \to \ul{\mf{u}} \to \ul{\mf{b}}^{\circ} \to \mc{O} \to 0
\end{displaymath}
of sheaves on $\mathbf{P}^1$. A global section of $\ul{\mf{b}}^{\circ}$ is determined by the endomorphism of $\mc{O}^2$
it induces, and this endomorphism has image in $\mc{O}(-1)$. Since there are no non-zero maps $\mc{O}^2 \to
\mc{O}(-1)$, we conclude that $\Gamma(\mathbf{P}^1, \ul{\mf{b}}^{\circ})=0$. This, combined with the above exact sequence,
implies that $\ul{\mf{b}}^{\circ} \cong \mc{O}(-1)^{\oplus 2}$.
\subsection{Strongly nilpotent matrices}
Let $T$ be a $k$-algebra. We say that a matrix $m$ in $M_2(T)$ is \emph{strongly nilpotent} if its trace and
determinant are both 0. This implies $m^2=0$, but the converse does not hold in general (it does if $T$ is a domain).
If $\mf{u}$ is a nilpotent subalgebra of $(M_2)_T$ then every element of $\mf{u}$ is strongly nilpotent.
\subsection{The space $\mc{A}$}
\label{aspace}
Let $r \ge 1$ be an integer. Let $\mc{A}=\mc{A}_r$ be the functor which assigns to a $k$-algebra $T$ the set
$\mc{A}(T)$ of tuples $(m_1, \ldots, m_r)$ where each $m_i \in M_2(T)$ is a strongly nilpotent matrix such that
$m_i m_j=0$ for all $i$ and $j$. We let $a \in \mc{A}(k)$ denote the tuple $(0, \ldots, 0)$. The main result of
this section is the following theorem:
\begin{theorem}
\label{thm-a}
The functor $\mc{A}$ is (represented by) a geometrically integral normal Cohen--Macaulay affine scheme of dimension
$r+1$. For $r>1$ the local ring at $a$ is not Gorenstein.
\end{theorem}
It is clear that $\mc{A}$ is an affine scheme. In fact, write
\begin{displaymath}
m_i=\mat{a_i}{b_i}{c_i}{-a_i}.
\end{displaymath}
Then $\mc{A}=\Spec(R)$, where $R$ is the quotient of $k[a_i, b_i, c_i]_{1 \le i \le r}$ by the equations
\begin{equation}
\label{a-eq}
a_ia_j=b_i c_j, \qquad a_i b_j=a_j b_i, \qquad a_i c_j=a_j c_i.
\end{equation}
For $i \ne j$, these equations express the identity $m_i m_j=0$. For $i=j$, the latter two equations are trivial,
while the first expresses that $m_i$ has determinant 0.
Let $\wt{\mc{A}}$ be the functor which attaches to a $k$-algebra $T$ the set of tuples $(\mf{u}; m_1, \ldots, m_r)$
where $\mf{u}$ is a nilpotent subalgebra of $(M_2)_T$ and each $m_i$ belongs to $\mf{u}$. It is clear that
$\wt{\mc{A}}$ is represented by the total space of the vector bundle $\ul{\mf{u}}^{\oplus r}$ over $\mathbf{P}^1$, and is
thus smooth and geometrically integral of dimension $r+1$. Let $\wt{R}$ be the ring of global functions on
$\wt{\mc{A}}$. Then $\wt{R}$ is normal and geometrically integral.
Let $(\mf{u}; m_1, \ldots, m_r)$ be an element of $\wt{\mc{A}}(T)$. Then each $m_i$ is strongly nilpotent, and
$m_i m_j=0$ for all $i$ and $j$. Thus $(m_1, \ldots, m_r)$ defines an element of $\mc{A}(T)$. We therefore have
a map of schemes $\wt{\mc{A}} \to \mc{A}$, and thus a corresponding map of rings $R \to \wt{R}$.
\begin{lemma}
The map $R \to \wt{R}$ is an isomorphism.
\end{lemma}
\begin{proof}
Let $\eta^{\vee}$ be the vector bundle $\ul{\mf{u}}^{\oplus r}$ on $\mathbf{P}^1$ and let $\epsilon^{\vee}$ be the bundle
$(\ul{\mf{g}}^{\circ})^{\oplus r}$. Then $\eta^{\vee}$ is naturally a subbundle of $\epsilon^{\vee}$; let
$\xi^{\vee}$ be the quotient bundle. We have isomorphisms $\eta=\mc{O}(2)^{\oplus r}$ and $\xi=\mc{O}(1)^{\oplus 2r}$.
Note that $\wt{R}=\Gamma(\mathbf{P}^1, \Sym(\eta))$. Let $S$ be the polynomial ring $k[a_i, b_i, c_i]$, which is
identified with $\Gamma(\mathbf{P}^1, \Sym(\epsilon))$.
We now apply Proposition~\ref{geo-2}. We find
\begin{displaymath}
\wt{R}/S_+ \wt{R}=H^0(\mathbf{P}^1, \mc{O}) \oplus H^1(\mathbf{P}^1, \xi)[1]=k,
\end{displaymath}
and so $S \to \wt{R}$ is surjective (by Nakayama's lemma). Let $I$ be the kernel. Then $\Tor^1_S(R, k)$ is
identified with $I/S_+ I$, and so Proposition~\ref{geo-2} gives
\begin{displaymath}
I/S_+ I=H^0(\mathbf{P}^1, \xi)[1] \oplus H^1(\mathbf{P}^1, \lw{2}{\xi})[2].
\end{displaymath}
The $H^0$ vanishes and the $H^1$ has dimension $\binom{r}{2}$. Thus $I$ is generated in degree 2 and its degree 2
piece has dimension $\binom{r}{2}$. An elementary argument shows that the $\binom{r}{2}$ quadratic elements of $S$
given in equation \eqref{a-eq} are linearly independent, and so the kernel of $S \to R$ is equal to $I$. Thus
the map $R \to \wt{R}$ is an isomorphism.
\end{proof}
The theorem now follows immediately from Proposition~\ref{geo-1}. To be precise, the above lemma shows that $R$
is normal and geometrically integral. Let $\eta$ as above. The space $Z$ in Proposition~\ref{geo-1} is just
$\wt{A}$, and the point 0 of $\Spec(R)$ is just $a$. The proposition shows that $\wt{\mc{A}} \to \mc{A}$ is an
isomorphism away from $a$, that $\wt{R}=R$ is Cohen--Macaulay and that the local ring of $R$ at $a$ is not Gorenstein
when $r>1$.
\begin{remark}
The ring $R$ is isomorphic to the projective coordinate ring of $\mathbf{P}^1 \times \mathbf{P}^r$ with respect to the bundle
$\mc{O}(2, 1)$. The singularities of such Segre--Veronese rings have been well-studied, and more general results
than Theorem~\ref{thm-a} appear in the literature; see, for instance, \cite[\S 0.4]{BarcanescuManolache}.
\end{remark}
\subsection{The space $\mc{B}$}
\label{bspace}
Let $\mc{B}=\mc{B}_r$ be the functor which assigns to a $k$-algebra $T$ the set $\mc{B}(T)$ of tuples $(\phi,
\alpha; m_1, \ldots, m_r)$ where $\phi$ is an element of $M_2(T)$ of determinant 1, $\alpha$ is an element of $T$
which is a root of the characteristic polynomial of $\phi$ and the $m_i$ are strongly nilpotent matrices in $M_2(T)$
such that $m_i m_j=0$ and $m_i \phi=\alpha m_i$. Let $b \in \mc{B}(k)$ be the point $(1, 1; 0, \ldots, 0)$. The
main result of this section is the following:
\begin{theorem}
\label{thm-b}
The functor $\mc{B}$ is (represented by) a geometrically integral normal Cohen--Macaulay affine scheme of dimension
$r+3$. The local ring at the point $b$ is not Gorenstein.
\end{theorem}
It is clear that $\mc{B}$ is an affine scheme. In fact, write
\begin{displaymath}
\phi=\mat{\phi_1}{\phi_2}{\phi_3}{\phi_4}
\end{displaymath}
and keep the notation for the $m_i$ from the previous section. Then $\mc{B}=\Spec(R)$, where $R$ is the quotient of
$k[a_i, b_i, c_i, \phi_j, \alpha]$ (with $1 \le i \le r$ and $1 \le j \le 4$) by the equations \eqref{a-eq}, the
equations
\begin{equation}
\label{b-eq-1}
a_i \phi_1+b_i \phi_3=\alpha a_i, \qquad
a_i \phi_2+b_i \phi_4=\alpha b_i, \qquad
c_i \phi_1-a_i \phi_3=\alpha c_i, \qquad
c_i \phi_2-a_i \phi_4=-\alpha a_i,
\end{equation}
the equation
\begin{equation}
\label{b-eq-2}
\alpha^2-(\phi_1+\phi_4) \alpha+(\phi_1 \phi_4-\phi_2 \phi_3)=0
\end{equation}
and the equation
\begin{equation}
\label{b-eq-3}
\phi_1 \phi_4-\phi_2 \phi_3=1.
\end{equation}
Of course, the equaion \eqref{b-eq-1} express the identity $m_i \phi=\alpha m_i$, while \eqref{b-eq-2} expresses that
$\alpha$ is a root of the characteristic polynomial of $\phi$ and \eqref{b-eq-3} expresses that $\phi$ has determinant
1.
Let $\mc{B}^{\circ}$ be defined like $\mc{B}$ except without any condition on the determinant of $\phi$. Then
$\mc{B}^{\circ}=\Spec(R^{\circ})$, where $R^{\circ}$ is the quotient of $k[a_i, b_i, c_i, \phi_j, \alpha]$ by the
equations \eqref{a-eq}, \eqref{b-eq-1} and \eqref{b-eq-2}. Note that these equations are all homogeneous of degree
two, and so $R^{\circ}$ is graded.
Let $\wt{\mc{B}}^{\circ}$ be the functor assigning to a $k$-algebra $T$ the set of tuples
$(\mf{b}; \phi; m_1, \ldots, m_r)$ where $\mf{b}$ is a Borel subalgebra of $\mf{g}_T$, $\phi$ is an element of $\mf{b}$
and the $m_i$ are elements of $\mf{u}$, the nilpotent radical of $\mf{b}$. It is clear that $\wt{\mc{B}}^{\circ}$
is represented by the total space of the vector bundle $\ul{\mf{b}} \oplus \ul{\mf{u}}^{\oplus r}$ over $\mathbf{P}^1$. We
write $\wt{R}^{\circ}$ for the ring of global functions on $\wt{B}^{\circ}$.
We let $\wt{\mc{B}}$ be defined like $\wt{\mc{B}}^{\circ}$ but with the condition $\det(\phi)=1$ imposed. Note
that $\wt{\mc{B}}$ is the fiber product of the universal Borel subgroup of $\mathrm{SL}(2)$ with a vector bundle over $\mathbf{P}^1$,
and is therefore smooth and geometrically integral of dimension $r+3$. We let $\wt{R}$ be the ring of global
functions on $\wt{\mc{B}}$.
Let $(\mf{b}; \phi; m_1, \ldots, m_r)$ be an element of $\wt{\mc{B}}^{\circ}(T)$. Let $\alpha$ be the eigenvalue
by which $\phi$ acts on the $[\mf{b}, \mf{b}]$ coinvariants of $T^2$ (e.g., the lower right entry of $\phi$ if $\mf{b}$
is upper triangular). Then $\alpha$ satisfies the characteristic
polynomial of $\phi$, and $m_i \phi=\alpha m_i$. It follows that $(\phi, \alpha; m_1, \ldots, m_r)$ is an element of
$\mc{B}^{\circ}$. We thus have a natural map $\wt{\mc{B}}^{\circ} \to \mc{B}^{\circ}$, and thus an induced map
$R^{\circ} \to \wt{R}^{\circ}$. Of course, we also have $\wt{\mc{B}} \to \mc{B}$ and $R \to \wt{R}$.
We summarize the above definitions with two commutative diagrams
\begin{displaymath}
\xymatrix{
\wt{\mc{B}} \ar@{^(->}[r] \ar[d] & \wt{\mc{B}}^{\circ} \ar[d] \\
\mc{B} \ar@{^(->}[r] & \mc{B}^{\circ} }
\qquad\qquad
\xymatrix{
\wt{R} & \wt{R}^{\circ} \ar[l] \\
R \ar[u] & R^{\circ} \ar@{->>}[l] \ar[u] }
\end{displaymath}
We now show that the vertical ring maps in the right diagram are isomorphisms.
\begin{lemma}
\label{lem-b1}
The map $R^{\circ} \to \wt{R}^{\circ}$ is an isomorphism.
\end{lemma}
\begin{proof}
Let $\eta^{\vee}$ be the vector bundle $\ul{\mf{b}} \oplus \ul{\mf{u}}^{\oplus r}$ on $\mathbf{P}^1$, which is isomorphic
to $\mc{O} \oplus \mc{O}(-1)^{\oplus (r+2)}$. The bundle $\eta^{\vee}$ is naturally
a subbundle of the constant bundle $\epsilon^{\vee}=\ul{\mf{g}} \oplus (\ul{\mf{g}}^{\circ})^{\oplus r}$.
The quotient bundle $\xi^{\vee}$ is isomorphic to $\mc{O}(2) \oplus \mc{O}(1)^{\oplus 2r}$. The ring $\wt{R}^{\circ}$
is $\Gamma(\mathbf{P}^1, \Sym(\eta))$. Let $S$ be the ring $\Gamma(\mathbf{P}^1, \Sym(\epsilon))$, i.e., the polynomial ring
$k[a_i, b_i, c_i, \phi_i]$. (Note: $S$ does not contain $\alpha$.) Then $\wt{R}^{\circ}$ is naturally an
$S$-algebra, as is $R^{\circ}$, and the map $R^{\circ} \to \wt{R}^{\circ}$ is $S$-linear.
It is clear that $R^{\circ}/S_+ R^{\circ}$ is isomorphic to $k \oplus k[1]$, and spanned by the images of 1 and
$\alpha$. Proposition~\ref{geo-2} shows that
\begin{displaymath}
\wt{R}^{\circ}/S_+ \wt{R}^{\circ}=H^0(\mathbf{P}^1, \mc{O}) \oplus H^1(\mathbf{P}^1, \xi)[1]=k \oplus k[1].
\end{displaymath}
Since $\alpha$ maps to a non-zero element of $\wt{R}^{\circ}$, we see that the map $R^{\circ} \to \wt{R}^{\circ}$
induces an isomorphism after quotienting by $S_+$, and is therefore surjective.
Consider the diagram
\begin{displaymath}
\xymatrix{
0 \ar[r] & M \ar[d] \ar[r] & S \oplus S[1] \ar[r] \ar@{=}[d] & R^{\circ} \ar[d] \ar[r] & 0 \\
0 \ar[r] & N \ar[r] & S \oplus S[1] \ar[r] & \wt{R}^{\circ} \ar[r] & 0 }
\end{displaymath}
The map $S \oplus S[1] \to R^{\circ}$ sends the basis vectors to 1 and $\alpha$. The map to $\wt{R}^{\circ}$ is
defined similarly. These maps are surjective by the previous paragraph. We let $M$ and $N$ denote their kernels,
so that $M \subset N$. Now, the equations for $\wt{R}^{\circ}$ are all of degree 2, which means that the natural
map $M_2 \to M/S_+M$ is an isomorphism. On the other hand,
$N/S_+N$ is identified with $\Tor_1^S(\wt{R}^{\circ}, k)$, and so another application of Proposition~\ref{geo-2}
yields
\begin{displaymath}
N/S_+ N=H^0(\mathbf{P}^1, \xi)[1] \oplus H^1(\mathbf{P}^1, \lw{2}{\xi})[2],
\end{displaymath}
and so $N/S_+ N$ is concentrated in degree 2 and of dimension $4r+\binom{2r}{2}$. We thus have a diagram
\begin{displaymath}
\xymatrix{
M_2 \ar[r] \ar[d] & M/S_+ M \ar[d] \\
N_2 \ar[r] & N/S_+ N }
\end{displaymath}
in which the horizontal maps are isomorphisms. Obviously, the left vertical map is injective. An elementary
argument shows the $\binom{2r}{2}$ equations in \eqref{a-eq} and the $4r$ equations in \eqref{b-eq-1} are linearly
independent, and so $M_2$ has dimension $4r+\binom{2r}{2}$. (Note: the equation \eqref{b-eq-2} does not appear in
$M$.) We therefore find that $M_2 \to N_2$ is surjective,
and so $M/S_+ M \to N/S_+ N$ is surjective, and so $M \to N$ is surjective, i.e., $M=N$. This completes the proof.
\end{proof}
\begin{lemma}
\label{lem-b3}
The map $R \to \wt{R}$ is an isomorphism.
\end{lemma}
\begin{proof}
Due to the previous lemma, it is enough to show that $\wt{R}^{\circ} \to \wt{R}$ is surjective with kernel generated
by $\det(\phi)-1$. Now, the space $\wt{\mc{B}}^{\circ}$ is affine over $\mathbf{P}^1$ and corresponds
to the sheaf of algebras $\Sym(\eta)$. The space $\wt{\mc{B}}$ is also affine over $\mathbf{P}^1$; let $\mc{R}$ denote the
corresponding sheaf of algebras. We have a short exact sequence of sheaves on $\mathbf{P}^1$
\begin{displaymath}
0 \to \Sym(\eta) \to \Sym(\eta) \to \mc{R} \to 0
\end{displaymath}
where the first map is given by multiplication by $\det(\phi)-1$, an element of $\Gamma(\mathbf{P}^1, \Sym(\eta))$. Since
$H^1(\mathbf{P}^1, \Sym(\eta))=0$, upon taking global sections we see that $\wt{R}^{\circ} \to \wt{R}$ is surjective and its
kernel is generated by $\det(\phi)-1$.
\end{proof}
Let $\mc{B}^{\circ}_0(T)$ be the subset of $\mc{B}^{\circ}(T)$ consisting of tuples $(\phi, \alpha; m_1, \ldots, m_r)$
where $m_i=0$ and $\phi=\alpha$ is a scalar matrix. Define $\wt{\mc{B}}^{\circ}$ similarly.
\begin{lemma}
\label{lem-b2}
The map $\wt{\mc{B}}^{\circ} \setminus \wt{\mc{B}}^{\circ}_0 \to \mc{B}^{\circ} \setminus \mc{B}^{\circ}_0$ is an
isomorphism. The space $\mc{B}^{\circ}$ is Cohen--Macaulay; moreover, it is smooth away from $\mc{B}^{\circ}_0$ and
not Gorenstein at any point in $\mc{B}^{\circ}_0$.
\end{lemma}
\begin{proof}
Let $\eta$ be as in Lemma~\ref{lem-b1}. The $\mc{O}$ summand of $\eta$ comes from the scalar matrices in $\mf{b}$.
Let $\eta'$ be the complement of this, so that $\eta=\eta' \oplus \mc{O}$. Put $A=\Gamma(\mathbf{P}^1, \Sym(\eta'))$. Then
$\wt{R}^{\circ}=R^{\circ}$ is the polynomial ring in one variable over $A$, and $\mc{B}^{\circ}=\Spec(A)
\times \mathbf{A}^1$, with $\{0 \} \times \mathbf{A}^1$ identified with $\mc{B}^{\circ}_0$. Proposition~\ref{geo-1}
applies to $\eta'$. We find that the map from the total space of $\eta'$ to $\Spec(A)$ is an isomorphism away from 0,
that $A$ is Cohen--Macaulay and that the local ring of $A$ at the point 0 is not Gorenstein. The lemma follows.
\end{proof}
\begin{lemma}
The map $\wt{\mc{B}} \to \mc{B}$ is birational.
\end{lemma}
\begin{proof}
This follows immediately from the previous lemma and the fact that $\mc{B}^{\circ}_0 \cap \mc{B}$ is a proper closed
set in $\mc{B}$ (in fact, it consists of two closed points).
\end{proof}
We now prove the theorem.
\begin{proof}[Proof of Theorem~\ref{thm-b}]
Since $\wt{\mc{B}}$ is normal and geometrically integral, so is $R=\wt{R}$. As $\wt{\mc{B}} \to \mc{B}$ is birational,
$R$ has dimension $r+3$.
Since $R$ is the quotient of the Cohen--Macaulay ring $R^{\circ}$ by the non-zerodivisor $\det(\phi)-1$, it is
Cohen--Macaulay \cite[Thm.~2.1.3a]{BrunsHerzog}. Furthermore, since $R^{\circ}$ is not Gorenstein at the point $b$, and
$\det(\phi)-1$ belongs to the maximal ideal at $b$, we see that $R$ is not Gorenstein at $b$
\cite[Prop.~3.1.19b]{BrunsHerzog}.
\end{proof}
\subsection{The space $\mc{C}$}
\label{cspace}
Let $\mc{C}=\mc{C}_r$ be the functor which assigns to a $k$-algebra $T$ the set $\mc{C}(T)$ of tuples
$(\phi, \alpha; m_1, \ldots, m_{r+1})$ where $\phi$ is an element of $M_2(T)$ of determinant 1, $\alpha$ is
an element of $T$ satisfying the characteristic polynomial of $\phi$ and the $m_i$ are strongly nilpotent matrices
in $M_2(T)$ such that the following equations hold:
\begin{displaymath}
m_i m_j=0, \qquad m_i \phi=\alpha m_i, \qquad m_{r+1} \phi=\phi m_{r+1}.
\end{displaymath}
The first two equations are for $1 \le i, j \le r+1$. The second equation is equivalent to $\phi m_i=\alpha^{-1} m_i$,
as can be seen by taking the adjugate of each side. In particular, the final equation is equivalent to
$(\alpha^2-1)m_{r+1}=0$. We let $c \in \mc{C}(k)$ denote the point
$(1, 1; 0, \ldots, 0)$. The main result of this section is the following theorem:
\begin{theorem}
\label{thm-c}
The functor $\mc{C}$ is (represented by) a geometrically reduced affine scheme which is equidimensional of dimension
$r+3$. It has three irreducible components: two isomorphic to $\mc{A}_{r+2}$ (defined by the equations
$\alpha=\pm 1$) and one isomorphic to $\mc{B}_r$ (defined by the equation $m_{r+1}=0$).
\end{theorem}
We require the following simple lemma.
\begin{lemma}
Let $R$ be a ring, let $\mf{p}$ be a prime ideal of $R$ and let $\mf{a}$ be a principal ideal of $R$ not contained in
$\mf{p}$. Then $\mf{a} \cap \mf{p}=\mf{a} \mf{p}$.
\end{lemma}
\begin{proof}
Clearly, $\mf{a} \mf{p}$ is contained in $\mf{a} \cap \mf{p}$. We now establish the reverse inclusion. Let
$\mf{a}=(a)$ and let $x$ be an element of $\mf{a} \cap \mf{p}$. We can then write $x=ay$ for some $y \in R$. Since
$ay$ belongs to $\mf{p}$ but $a$ does not belong to $\mf{p}$, we conclude that $y$ belongs to $\mf{p}$. Thus $x$
belongs to $\mf{a} \mf{p}$.
\end{proof}
We now prove the theorem.
\begin{proof}[Proof of Theorem~\ref{thm-c}]
It is clear that $\mc{C}$ is represented by an affine scheme $\Spec(R)$; we do not write the equations, but keep our
previous notation for elements of the ring $R$. The locus $\alpha=\pm 1$ in
$\mc{C}$ is isomorphic to $\mc{A}_{r+2}$, the isomorphism taking a $T$-point $(\phi, \alpha; m_1, \ldots, m_{r+1})$ to
$(\phi-\alpha, m_1, \ldots, m_{r+1})$. The locus $m_{r+1}=0$ in $\mc{C}$ is isomorphic to $\mc{B}_r$ (obviously). It
follows that $\mf{p}_1=(\alpha-1)$, $\mf{p}_2=(\alpha+1)$ and $\mf{p}_3=(a_{i+1}, b_{i+1}, c_{i+1})$ are prime ideals
of $R$. We claim that their intersection is the zero ideal; this will prove the theorem. Since the rings
$R/\mf{p}_i$ all have the same dimension, there is no containment between the primes $\mf{p}_i$. Since
$\mf{p}_1$ and $\mf{p}_2$ are both prime principal ideals, the lemma gives $\mf{p}_1 \cap \mf{p}_2=\mf{p}_1 \mf{p}_2$,
which is again a principal ideal. A second application of the lemma gives $\mf{p}_1 \cap \mf{p}_2 \cap \mf{p}_3=
\mf{p}_1 \mf{p}_2 \mf{p}_3$. However, since $(\alpha^2-1)m_{r+1}=0$, the product of the $\mf{p}_i$ is zero. This
completes the proof.
\end{proof}
\begin{remark}
The point $c$ belongs to each of the three irreducible components of $\mc{C}$, and corresponds to the points $a$
and $b$ when these components are identified with $\mc{A}$ and $\mc{B}$.
\end{remark}
\section{Local deformation rings}
\label{s:local}
In this section we study certain ordinary deformation rings of local Galois groups.
\subsection{Set-up}
Let $F$ be a finite extension of $\mathbf{Q}_p$ of degree $d$ with absolute Galois group $G_F$. Let $E$ be a finite extension
of $\mathbf{Q}_p$ with ring of integers $\mc{O}$, uniformizer $\pi$ and residue field $k$. Let $V_0$ be a two dimensional
$k$-vector space
equipped with a continuous action of $G_F$ having cyclotomic determinant. Let $\ms{C}_{\mc{O}}$ denote the category of
artinnian local $\mc{O}$-algebras with residue field $k$. Define a functor\footnote{The appropriate 2-categorical
notions are to be understood throughout.} $D$ on $\ms{C}_{\mc{O}}$ by assigning to an algebra $A$ the groupoid of all
deformations of $V_0$ to $A$ with cyclotomic determinant.
That is, $D(A)$ is the category whose objects are free rank two $A$-modules $V$ equipped with a continuous action of
$G_F$ (for the discrete topology) having cyclotomic determinant together with an isomorphism $V \otimes_A k \to V_0$;
the morphisms in $D(A)$ are isomorphisms.
The category $D(A)$ is typically not discrete, and so $D$ will typically not be representable. To circumvent this
annoyance,
we use framed deformations. For $A$ as above, define $D^{\Box}(A)$ to be the category of pairs $(V, \{e_1, e_2\})$,
where $V$ is an object of $D(A)$ and $\{e_1, e_2\}$ is a basis of $V$ as an $A$-module. Morphisms in $D^{\Box}(A)$
are required to respect the basis, and so there is at most one morphism between two objects. The functor
$D^{\Box}(A)$ is pro-representable by a complete local noetherian $\mc{O}$-algebra $R^{\mathrm{univ}}$ together with a
free rank two $R^{\mathrm{univ}}$-module $V^{\mathrm{univ}}$. The action of $G_F$ on $V^{\mathrm{univ}}$ is continuous when $V^{\mathrm{univ}}$
is given the $\mf{m}^{\mathrm{univ}}$-adic topology ($\mf{m}^{\mathrm{univ}}$ being the maximal ideal of $R^{\mathrm{univ}}$).
\subsection{Statement of problem}
For the purpose of this paper, we say that a continuous representation $\rho:G_F \to \mathrm{GL}_2(E')$, with $E'$ a finite
extension of $E$, is \emph{ordinary} if
\begin{displaymath}
\rho \vert_{I_F} \cong \mat{\chi}{\ast}{}{1}.
\end{displaymath}
Let $X$ be the locus in $\MaxSpec(R^{\mathrm{univ}}[1/\pi])$ consisting of ordinary representations. Kisin's method,
discussed briefly below, shows
that $X$ is Zariski closed, and so there is a unique $\mc{O}$-flat reduced quotient $R$ of $R^{\mathrm{univ}}$ such that
$\MaxSpec(R[1/\pi])=X$. This method also allows one to compute the components of $R[1/\pi]$ and show that each of
them is formally smooth over $E$. However, as discussed in the introduction, the method does not provide much
information about $R$ itself. Our goal is to understand this ring as best we can.
A natural way to proceed is to try to understand the deformation problem $R$ represents. A good first guess is
that giving a map $R \to A$ is the same as giving a deformation $V$ of $V_0$ to $A$ which is in some sense
``ordinary.'' To this end, let us call such a $V$ \emph{ordinary} if there exists a rank one $A$-module summand $L$
of $V$ on which $G_F$ acts through the cyclotomic character and such that $G_F$ acts on $V/L$ trivially. Let
$\ol{D}^{\mathrm{ord}}$ denote the functor assigning to $A$ the groupoid of ordinary deformations.
If $V_0$ is non-trivial then the functor $\ol{D}^{\mathrm{ord}, \Box}$ is represented by $R$. This gives a moduli-theoretic
description of $R$, and allows $R$ to be studied relatively easily. For instance, one can show
that the irreducible components of $\Spec(R)$ are formally smooth over $\mc{O}$ --- in fact, this comes out
of Kisin's analysis.
If $V_0$ is trivial then the functor $\ol{D}^{\mathrm{ord}, \Box}$ is not representable. The problem is that, in this case,
a line $L$ in $V$ as above need not be unique. The main idea of Kisin's method is to study the functor $D^{\mathrm{ord}}$
parameterizing pairs $(V, L)$, with $L$ as above. The framed version of this functor is representable, but not by $R$;
rather, $\Spec(R)$ is the
scheme-theoretic image of the representing object in $\Spec(R^{\mathrm{univ}})$. It is hard to deduce properites of $R$
from this description. In the rest of \S \ref{s:local}, we will follow the plan outlined in the introduction to
obtain the equations cutting $R$ out from $R^{\mathrm{univ}}$, and then use this to analyze $R$. (Actually, we work with a
slightly modified ring $\wt{R}$ defined below.)
{\it We assume for the rest of \S \ref{s:local} that $V_0$ is trivial.}
\subsection{Three sets of Galois representations}
Let $x$ be a point in $\MaxSpec(R^{\mathrm{univ}}[1/\pi])$ with residue field $E_x$. Then $E_x$ is a finite extension of $E$,
and the action of $G_F$ on $V_x=V^{\mathrm{univ}} \otimes_{R^{\mathrm{univ}}} E_x$ is continuous when $V_x$ is given its
$p$-adic topology.
As defined above, we let $X \subset \MaxSpec(R^{\mathrm{univ}}[1/\pi])$ denote the set of points $x$ for which $V_x$ is
ordinary. We now define three subsets of $X$. We let $X_1$ be the subset consisting of points $x$ where $V_x$
is an extension of $E_x$ by $E_x(\chi)$ on the full Galois group. Similarly, we let
$X_2$ denote the subset consisting of points $x$ where $V_x$ is an extension of $E_x(\eta)$ by
$E_x(\eta \chi)$ on the full Galois group, where $\eta$ is the unramified quadratic character. Finally, we let
$X_3$ denote the subset of $X$ where the representation is crystalline. We have the following basic result.
\begin{lemma}
The set $X$ is the union of the subsets $X_1$, $X_2$ and $X_3$.
\end{lemma}
\begin{proof}
Let $x \in X$. Then $V_x \vert_{I_F}$ is an extension of $E_x$ by $E_x(\chi)$, and so $V_x$ is an extension of
$E_x(\psi^{-1})$ by $E_x(\psi \chi)$ for some unramified character $\psi$. Thus $V_x$ defines an element of
$H^1(G_F, E_x(\chi \psi^2))$.
Now, we have isomorphisms
\begin{displaymath}
H^1(G_F, E_x(\chi \psi^2)) \cong H^1(I_F, E_x(\chi \psi^2))^{\Gal(F^{\mathrm{un}}/F)} = (((F^{\mathrm{un}})^{\times})^{\wedge} \otimes
E_x(\psi^2))^{\Gal(F^{\mathrm{un}}/F)}.
\end{displaymath}
The first isomorphism is restriction, the second comes from Kummer theory. Here the $\wedge$ denotes $p$-adic
completion. The valuation map on $(F^{\mathrm{un}})^{\times}$ defines a map
\begin{displaymath}
H^1(G_F, E_x(\chi \psi^2)) \to E_x(\psi^2)^{\Gal(F^{\mathrm{un}}/F)}.
\end{displaymath}
We thus see that either $\psi^2$ is trivial, in which case $x$ belongs to $X_1$ or $X_2$, or else the above
map is zero, in which case $x$ belongs to $X_3$.
\end{proof}
\subsection{The ring $R$}
The following result is due to Kisin.
\begin{proposition}[Kisin]
\label{local-1}
Let $\ast \in \{1, 2, 3\}$. Then there exists a unique reduced $\mc{O}$-flat quotient $R_{\ast}$ of $R^{\mathrm{univ}}$
such that $\MaxSpec(R_{\ast}[1/\pi])$ is equal to $X_{\ast}$. The ring $R_{\ast}$ is a domain which is
equidimensional of dimension $d+4$. The ring $R_{\ast}[1/\pi]$ is regular.
\end{proposition}
We recall the relevant pieces of the argument. For a complete proof, see \S 2.4 of Kisin's paper \cite{Kisin2}
or Conrad's unpublished notes \cite{Conrad}. We just deal with the $\ast=3$ case; the other cases are similar. Define
a functor $D_3$ on the category $\ms{C}_{\mc{O}}$ by assigning to an algebra $A$ the groupoid $D_3(A)$ of
pairs $(V, L)$ where $V \in D(A)$ and $L$ is a rank one $A$-module summand of $V$ on which $G_F$ acts through $\chi$
and such that the class in $H^1(G_F, L \otimes (V/L)^{\vee})$ determined by the extension
\begin{displaymath}
0 \to L \to V^{\mathrm{univ}}_A \to V^{\mathrm{univ}}_A/L \to 0
\end{displaymath}
belongs to $H^1_f(G_F, L \otimes (V^{\mathrm{univ}}_A/L)^{\vee})$ (see \cite[\S 2.4.1]{Kisin2} for the definition of $H^1_f$).
There is a natural
map $\Theta:D_3 \to D$ which forgets $L$. This map is relatively representable and
projective; in fact, there is a closed immersion $D_3 \to \mathbf{P}^1_D$ lifting $\Theta$. It follows that
$D_3^{\Box}$ is representable by a projective formal scheme $\wh{Z}$ over $\Spec(R^{\mathrm{univ}})$.
We let $Z$ be the algebraization of $\wh{Z}$, and still write $\Theta$ for the map $Z \to \Spec(R^{\mathrm{univ}})$.
The scheme $Z$ is formally smooth over $\mc{O}$.
The map $\Theta[1/\pi]$ is a closed immersion and induces a bijection between the closed points of $Z[1/\pi]$ and
the set $X_3 \subset \MaxSpec(R^{\mathrm{univ}}[1/\pi])$. It follows that if we let $R_3$ be such that $\Spec(R_3)$ is the
scheme-theoretic image of $\Theta$, then $R_3$ is $\mc{O}$-flat and reduced and satisfies $\MaxSpec(R_3[1/\pi])=X_3$.
It is clear that $R_3$ is the unique quotient of $R^{\mathrm{univ}}$ with these properties. Since $Z$ is formally smooth over
$\mc{O}$ and
$\Theta:Z[1/\pi] \to \Spec(R_3[1/\pi])$ is an isomorphism, it follows that $R_3[1/\pi]$ is formally smooth over
$E$, and thus regular. One deduces that $R_3$ is a domain from the fact that the fiber of $\Theta$
over the closed point of $R^{\mathrm{univ}}$ is connected (it is $\mathbf{P}^1$ since $V_0$ is trivial); see \cite[Cor.~2.4.6]{Kisin}.
The dimension of $R_3$ can be calculated by looking at a tangent space.
From Proposition~\ref{local-1}, we immediately obtain the following theorem:
\begin{proposition}
There exists a unique reduced $\mc{O}$-flat quotient $R$ of $R^{\mathrm{univ}}$ with the property that $\MaxSpec(R[1/\pi])$
is equal to $X$. The ring $R$ is equidimensional of dimension $d+4$ and has three minimal primes, namely the kernels
of the surjections $R \to R_{\ast}$.
\end{proposition}
\subsection{The ring $\wt{R}$}
\label{ss:rtilde}
Fix a Frobenius element $\phi$ of $G_F$. We assume $\chi(\phi)=1$ for convenience. Let $\wt{D}$ be the functor
on $\ms{C}_{\mc{O}}$ assigning to $A$ the set of
pairs $(V, \alpha)$, where $V$ belongs to $D(A)$ and $\alpha \in A$ is a root of the characteristic polynomial of
$\phi$ on $V$. The framed version $\wt{D}^{\Box}=\wt{D} \times_D D^{\Box}$ is pro-representable by a complete
local noetherian $\mc{O}$-algebra $\wt{R}^{\mathrm{univ}}$. The ring $\wt{R}^{\mathrm{univ}}$ is the quotient of $R^{\mathrm{univ}}[\alpha]$
by a monic degree two polynomial (the characteristic polynomial of $\phi$ on $V^{\mathrm{univ}}$).
We now define a map
\begin{displaymath}
\Phi:X \to \MaxSpec(\wt{R}^{\mathrm{univ}}[1/\pi]).
\end{displaymath}
Thus let $x$ be a point in $X$. The space $V_x$ contains a unique line $L_x$ on which $I_F$ acts through the
cyclotomic character. This line is stable by $G_F$ since $I_F$ is a normal subgroup. Let $\alpha_x$ be the scalar
through which $\phi$ acts on $V_x/L_x$. Then $\wt{x}=(V_x, \alpha_x)$
defines a point of $\MaxSpec(\wt{R}^{\mathrm{univ}})$. We put $\Phi(x)=\wt{x}$. Put $\wt{X}=\Phi(X)$ and
$\wt{X}_{\ast}=\Phi(X_{\ast})$.
The main result of this section is the following:
\begin{proposition}
Let $\ast \in \{1,2,3\}$. Then there is a unique reduced $\mc{O}$-flat quotient $\wt{R}_{\ast}$ of $\wt{R}^{\mathrm{univ}}$
such that $\MaxSpec(\wt{R}_{\ast}[1/\pi])$ is equal to $\wt{X}_{\ast}$. The ring $\wt{R}_{\ast}$ is a domain which is
equidimensional of dimension $d+4$. The natural map $R_{\ast} \to \wt{R}_{\ast}$ is an isomorphism after
inverting $\pi$.
\end{proposition}
This proposition can be proved in the same manner as Proposition~\ref{local-1}. Instead of doing this, we deduce it
from Proposition~\ref{local-1}, as this is a bit shorter. We only treat the $\ast=3$ case, as the others are similar.
Let $Z$ be the scheme constructed in the previous section. Let $\wt{\Theta}:Z \to \Spec(\wt{R}^{\mathrm{univ}})$ be the map
defined by taking $(V, L)$ to $(V, \alpha)$, where $\alpha$ is the scalar through which $\phi$ acts on $V/L$. We
have a commutative diagrams
\begin{displaymath}
\xymatrix{
& Z \ar[ld]_{\wt{\Theta}} \ar[rd]^{\Theta} \\
\Spec(\wt{R}^{\mathrm{univ}}) \ar[rr] && \Spec(R^{\mathrm{univ}}) }
\end{displaymath}
where the bottom horizontal map is the natural one (forget $\alpha$), and
\begin{displaymath}
\xymatrix{
& Z[1/\pi]' \ar[ld]_{\wt{\Theta}} \ar[rd]^{\Theta} \\
\wt{X}_3 && X_3 \ar[ll]_{\Phi} }
\end{displaymath}
where the prime denotes the set of closed points. Since $\Theta[1/p]$ is a closed immersion, it follows from the
first diagram that $\wt{\Theta}[1/p]$ is as well. We thus see that $\wt{\Theta}$ is injective in the second diagram;
since we know that $\Theta$ and $\Phi$ and bijections, it follows that $\wt{\Theta}$ is as well.
Let $\wt{R}_3$ be such that $\Spec(\wt{R}_3)$ is the scheme-theoretic
image of $\wt{\Theta}$. Since $Z$ is formally smooth over $\mc{O}$, the ring $\wt{R}_3$ is $\mc{O}$-flat
and reduced; furthermore, by the above comments, $\MaxSpec(\wt{R}_3[1/\pi])=\wt{X}_3$. It is clear that $\wt{R}_3$ is
the
unique quotient of $\wt{R}^{\mathrm{univ}}$ with these properties. Since $\Theta[1/\pi]:Z[1/\pi] \to \Spec(R_3[1/\pi])$ and
$\wt{\Theta}[1/\pi]:Z[1/\pi] \to \Spec(\wt{R}_3[1/\pi])$ are both isomorphisms, it follows from the first diagram
above that $R_3[1/\pi] \to \wt{R}_3[1/\pi]$ is an isomorphism. We thus conclude that $\wt{R}_3$ is a domain and
equidimensional of dimension $d+4$ from the corresponding results for $R_3$.
\begin{remark}
The map $R_{\ast} \to \wt{R}_{\ast}$ is an isomorphism for $\ast \in \{1,2\}$.
\end{remark}
The above proposition immediately implies the following one.
\begin{proposition}
\label{local-2}
There is a unique reduced $\mc{O}$-flat quotient $\wt{R}$ of $\wt{R}^{\mathrm{univ}}$ such that $\MaxSpec(\wt{R}[1/\pi])$ is
equal to $\wt{X}$. The ring $\wt{R}$ is equidimensional of dimension $d+4$ and has three minimal primes, the kernels
of the surjections $\wt{R} \to \wt{R}_{\ast}$. The natural map $R \to \wt{R}$ is an isomorphism after inverting $\pi$.
\end{proposition}
\subsection{The ring $\wt{R}^{\dag}$}
\label{ss:rdag}
For $A \in \ms{C}_{\mc{O}}$, let $\wt{D}^{\dag}(A)$ denote the subset of $\wt{D}(A)$ consisting of those pairs
$(V, \alpha)$ such that the following conditions are satisfied:
\begin{itemize}
\item $\tr{g}=\chi(g)+1$ for all $g \in I_F$.
\item $(g-1)(g'-1)=(\chi(g)-1)(g'-1)$ for $g,g' \in I_F$.
\item $(g-1)(\phi-\alpha)=(\chi(g)-1)(\phi-\alpha)$ for $g \in I_F$.
\item $(\phi-\alpha)(g-1)=(\alpha^{-1}-\alpha)(g-1)$ for $g \in I_F$.
\end{itemize}
The functor $\wt{D}^{\dag, \Box}$ is clearly prorepresentable by a ring $\wt{R}^{\dag}$; to obtain $\wt{R}^{\dag}$,
simply form the quotient of $\wt{R}^{\mathrm{univ}}$ by the above equations.
\begin{lemma}
The natural map $\wt{R}^{\mathrm{univ}} \to \wt{R}$ factors through $\wt{R}^{\dag}$.
\end{lemma}
\begin{proof}
The map $\wt{\Theta}:Z \to \Spec(\wt{R}^{\mathrm{univ}})$ clearly factors through the closed immersion $\Spec(\wt{R}^{\dag})
\to \Spec(\wt{R}^{\mathrm{univ}})$, which proves the
lemma.
\end{proof}
\subsection{The main theorems}
We prove two main theorems. The first is the following:
\begin{theorem}
\label{thm1}
The natural map $\wt{R}^{\dag} \to \wt{R}$ is an isomorphism.
\end{theorem}
This theorem gives a description of the points of $\wt{R}$. Our second theorem is the following:
\begin{theorem}
\label{thm2}
Let $\ast \in \{1, 2, 3\}$. The ring $\wt{R}_{\ast}$ is normal and Cohen--Macaulay but not Gorenstein.
\end{theorem}
The rest of this section is devoted to proving these two theorems. We begin with some lemmas.
\begin{lemma}
\label{lem-a}
The natural map $\wt{R}^{\dag}[1/\pi]_{\mathrm{red}} \to \wt{R}[1/\pi]$ is an isomorphism.
\end{lemma}
\begin{proof}
Let $\wt{x}$ be a point in $\MaxSpec(\wt{R}^{\dag}[1/\pi])$ and let $(V, \alpha)$ be the corresponding representation
and eigenvalue of $\phi$. Let $x$ be the image of $\wt{x}$ in $\MaxSpec(R^{\mathrm{univ}}[1/\pi])$, so that $V=V_x$. The
equations of \S \ref{ss:rdag} hold on $V$. The first of these, namely
$\tr(g)=\chi(g)+1$ for $g \in I_F$, shows that the semi-simplification of $V \vert_{I_F}$ is $\chi \oplus 1$.
Suppose that $V \vert_{I_F}$ is an extension of $\chi$ by 1, so that with respect to a suitable basis the action
of $I_F$ is given by
\begin{displaymath}
g \mapsto \mat{1}{f(g)}{}{\chi(g)}.
\end{displaymath}
The second equation of \S \ref{ss:rdag} shows that
\begin{displaymath}
(\chi(g)-1) f(g')=(\chi(g')-1) f(g)
\end{displaymath}
for all $g, g' \in I_F$, which shows that the cocycle $f$ is a coboundary (fix $g'$ with $\chi(g') \ne 1$). Thus
the extension is split.
The previous paragraph shows that we can regard $V \vert_{I_F}$ as an extension of 1 by $\chi$. Thus $x$ belongs
to $X$. Let $\beta$ be the eigenvalue of $\phi$ on the inertial coinvariants of $V$.
Then the fourth equation in \S \ref{ss:rdag} shows that $(\beta^{-1}-\alpha) (g-1)=(\alpha^{-1}-\alpha) (g-1)$ holds on
$V_x$ for all $g \in I_F$. This implies $\beta=\alpha$ (consider $g \in I_F$ with $\chi(g) \ne 1$), and so
$\wt{x}=\Phi(x)$. This shows that $\wt{x}$ belongs to $\wt{X}$.
We have just shown that the inclusion $\wt{X}=\MaxSpec(\wt{R}[1/\pi]) \subset \MaxSpec(\wt{R}^{\dag}[1/\pi])$ induced
by the surjection $\wt{R}^{\dag} \to \wt{R}$ is in fact an equality. The lemma follows.
\end{proof}
\begin{lemma}
\label{lem-b}
The ring $\wt{R}^{\dag}/\pi \wt{R}^{\dag}$ is isomorphic to the complete local ring of the scheme $\mc{C}_d$
at the point $c$ (see \S \ref{cspace} for the definition of $\mc{C}_d$ and $c$, and recall $d=[F:\mathbf{Q}_p]$).
\end{lemma}
\begin{proof}
Let $\wh{\mc{C}}$ denote the formal completion of $\mc{C}_d$ at the point $c$. Let $\ms{C}_k$ denote the category of
complete local noetherian $k$-algebras with residue field $k$. For $A \in
\ms{C}_k$, the set $\wh{\mc{C}}(A)$ consists of those elements of $\mc{C}(A)$ whose image in $\mc{C}(k)$ is the
point $c$. We will show that the functors $\wt{D}^{\dag, \Box}$ and $\wh{\mc{C}}$ are isomorphic on the category
$\ms{C}_k$. As $\wt{R}^{\dag, \Box}/\pi \wt{R}^{\dag}$ represents the former
functor, this will prove the lemma.
Let $A \in \ms{C}_k$. We regard elements of $\wt{D}^{\Box}(A)$ as pairs
$(\rho, \alpha)$ where $\rho:G_F \to \mathrm{GL}_2(A)$ is a homomorphism reducing to the trivial homomorphism modulo the
maximal ideal of $A$ and $\alpha$ is an element of $A$ satisfying the characteristic polynomial of $\rho(\phi)$.
An element $(\rho, \alpha)$ of $\wt{D}^{\Box}(A)$ belongs to $\wt{D}^{\dag, \Box}(A)$ if and only if the following
equations hold:
\begin{itemize}
\item $\tr{\rho(g)}=2$ for all $g \in I_F$.
\item $(\rho(g)-1)(\rho(g')-1)=0$ for $g, g' \in I_F$.
\item $(\rho(g)-1)(\rho(\phi)-\alpha)=0$ for $g \in I_F$.
\item $(\rho(\phi)-\alpha)(\rho(g)-1)=(\alpha-\alpha^{-1})(\rho(g)-1)$ for $g \in I_F$.
\end{itemize}
These equations come from combining the defining equations of $\wt{D}^{\dag}$ with the assumption that $\chi$ reduces
to 1 modulo $p$. Of course, we also have $\det{\rho(g)}=1$ for any such deformation. These conditions imply that
$\rho(g)-1$ is strongly nilpotent for any $g \in I_F$, and thus $\rho(g)^p=1$ for any such $g$. We thus see that if
$(\rho, \alpha)$ belongs to $\wt{D}^{\dag, \Box}(A)$ then $\rho \vert_{I_F}$ factors through the maximal abelian
quotient of $I_F$ of exponent $p$. Of course, $\rho$ factors through the maximal $p$-power quotient of $G_F$ since its
reduction modulo the maximal ideal of $A$ is trivial.
Let $G$ be the maximal $p$-power quotient of $G_F$ in which inertia is abelian and of exponent $p$. Let $U$ be the
inertia group in $G$. We have a short exact sequence
\begin{displaymath}
0 \to U \to G \to \mathbf{Z}_p \to 0.
\end{displaymath}
The image of $\phi$ is a topological generator of $\mathbf{Z}_p$. We give $U$ the structure of an $\mathbf{F}_p\lbb T \rbb$-module
by letting $T$ act by $\phi-1$. As computed in Proposition~\ref{galstruct}, $U$ is isomorphic to
$\mathbf{F}_p \oplus \mathbf{F}_p \lbb T \rbb^{\oplus d}$. Let $g_1, \ldots, g_d$ be an $\mathbf{F}_p \lbb T \rbb$-basis for the free
part of $U$ and let $g_{d+1}$ be a generator of the $T$-torsion of $U$. Note that to give a continuous map from $G$ to
some discrete group $\Gamma$ is the same as giving elements $\ol{\phi}$ and $\ol{g}_1, \ldots, \ol{g}_{d+1}$ of
$\Gamma$ such that the $\ol{g}_i$ commute with each other, $\ol{g}_i^p=1$ for each $i$, $\ol{\phi}$ has finite order
and $\ol{\phi}$ and $\ol{g}_{d+1}$ commute.
For $A \in \ms{C}_k$, we define a map $\wt{D}^{\dag, \Box}(A) \to \wh{\mc{C}}(A)$ by taking $(\rho, \alpha)$ to
the tuple $(\phi, \alpha; m_1, \ldots, m_{d+1})$ where $\phi=\rho(\phi)$ (apologies for the bad notation) and
$m_i=\rho(g_i)-1$. The defining equations for $\wt{D}^{\dag, \Box}$ given above show that this map is a
bijection.
\end{proof}
We now prove the first theorem.
\begin{proof}[Proof of Theorem~\ref{thm1}]
We follow the plan laid out in \S \ref{ss:out}. We have already completed step (a) by guessing
the equations for $\wt{R}$ and defining the ring $\wt{R}^{\dag}$. We now complete the process.
\begin{enumerate}
\setcounter{enumi}{1}
\item By Lemma~\ref{lem-a}, $\wt{R}^{\dag}[1/\pi]_{\mathrm{red}} \to \wt{R}[1/\pi]$ is an isomorphism. In particular,
by Proposition~\ref{local-2},
$\wt{R}^{\dag}[1/\pi]$ is equidimensional of dimension $d+3$ and has three minimal primes.
\item By Lemma~\ref{lem-b} and Theorem~\ref{thm-c}, $\wt{R}^{\dag}/\pi \wt{R}^{\dag}$ is reduced, equidimensional
of dimension $d+3$ and has three minimal primes.
\item By Proposition~\ref{flat}, the ring $\wt{R}^{\dag}$ is flat over $\mc{O}$. By Proposition~\ref{reduced} it is
reduced.
\item Since $\wt{R}^{\dag}$ is $\mc{O}$-flat and reduced and the map $\wt{R}^{\dag}[1/\pi]_{\mathrm{red}} \to \wt{R}[1/\pi]$
is an isomorphism, it follows that $\wt{R}^{\dag} \to \wt{R}$ is an isomorphism.
\end{enumerate}
This completes the proof.
\end{proof}
We now turn to the second theorem.
\begin{proof}[Proof of Theorem~\ref{thm2}]
As shown in the proof of Proposition~\ref{flat}, the ring $\wt{R}$ has three minimal primes, and these minimal primes
are naturally in correspondence with those of $\wt{R}/\pi \wt{R}$ and $\wt{R}[1/\pi]$. It is clear that two of
these minimal primes are defined by the equations $\alpha=1$ and $\alpha=-1$. The quotients by these minimal primes
are the rings $\wt{R}_1$ and $\wt{R}_2$. We thus see that $\wt{R}_1/\pi \wt{R}_1$ and $\wt{R}_2/\pi \wt{R}_2$ are both
isomorphic to the complete local ring of $\mc{A}_{d+2}$ at $a$. Finally, the third minimal prime gives $\wt{R}_3$. It
is clear that $\wt{R}_3/\pi \wt{R}_3$ is isomorphic to the complete local ring of $\mc{B}_d$ at $b$, since this is the
only thing left over.
Appealing to Theorem~\ref{thm-a} and Theorem~\ref{thm-b}, we see that $\wt{R}_{\ast}/\pi \wt{R}_{\ast}$ is integral,
normal, Cohen--Macaulay and not Gorenstein. The ring $R_{\ast}$ is a domain and $R_{\ast}[1/\pi]$ is normal (as it is
regular). It follows that $R_{\ast}$ is normal (by Proposition~\ref{normal}), Cohen--Macaulay
(by \cite[Thm.~2.1.3a]{BrunsHerzog}) and not Gorenstein (by \cite[Prop.~3.1.19b]{BrunsHerzog}).
\end{proof}
\begin{remark}
We can analyze the rings $\wt{R}_1$ and $\wt{R}_2$ directly, without using the ring $\wt{R}$. Indeed, we can
define $\wt{R}_1$ as the quotient of $\wt{R}$ be the equation $\alpha=1$, which realizes it directly as a quotient of
$R^{\mathrm{univ}}$. We can then go through the above arguments, but specifically for $\wt{R}_1$. However, we have not found
a way to analyze $\wt{R}_3$ directly: we do not know the equations that cut it out from $\wt{R}$. Note, however,
that we do know how to cut out $\wt{R}_3/\pi \wt{R}_3$ from $\wt{R}/\pi \wt{R}$: it is defined by the equation
$\rho(g_{d+1})=1$, where $g_{d+1}$ is as in the proof of Lemma~\ref{lem-b}. We have therefore studied $\wt{R}_3$
indirectly by studying the entire ring $\wt{R}$.
\end{remark}
\section{Global deformation rings}
\label{s:global}
In this section we give some global applications of the local results of the previous section. These applications
are really just some superficial remarks, and reasonably well-known, so we do not bother going into many details.
\subsection{Torsion-freeness of deformation rings}
Let $E$, $\mc{O}$ and $k$ be as in the previous section. Let $F$ be a totally real field, let $\Sigma$ be a finite
set of finite places of $F$, including all those above $p$, and let $\ol{\rho}:G_{F, \Sigma} \to \mathrm{GL}_2(k)$ be a
totally odd continuous representation of the absolute Galois group of $F$ unramified away from $\Sigma$. We assume
that $\ol{\rho}$ is absolutely irreducible and has determinant $\chi$, the cyclotomic character. We define several
deformation rings (all with fixed determinant $\chi$):
\begin{itemize}
\item For $v \in \Sigma$, let $R_v^{\Box, \mathrm{univ}}$ denote the universal framed deformation ring of
$\ol{\rho} \vert_{G_{F_v}}$.
\item For $v \in \Sigma$, choose a finite $R_v^{\Box, \mathrm{univ}}$-algebra $R_v^{\Box}$ which $\mc{O}$-flat and
equidimensional of dimension $[F_v:\mathbf{Q}_p]+4$ if $p \mid v$ or 4 if $p \nmid v$.
\item Let $R^{\Box, \mathrm{univ}}_{\mathrm{loc}}$ be the completed tensor product of the $R_v^{\Box, \mathrm{univ}}$ over $\mc{O}$ and let
$R^{\Box}_{\mathrm{loc}}$ be the completed tensor product of the $R_v^{\Box}$ over $\mc{O}$.
\item Let $R^{\Box, \mathrm{univ}}$ be the universal deformation ring of $\ol{\rho}$ with framings at each $v \in \Sigma$ and
let $R^{\Box}$ be the completed tensor product of $R^{\Box, \mathrm{univ}}$ with $R^{\Box}_{\mathrm{loc}}$ over
$R^{\Box, \mathrm{univ}}_{\mathrm{loc}}$.
\item Let $R^{\mathrm{univ}}$ be the universal (unframed) deformation ring of $\ol{\rho}$ and let $R$ be the descent of
$R^{\Box}$ from $R^{\Box, \mathrm{univ}}$ to $R^{\mathrm{univ}}$; it is a finite $R^{\mathrm{univ}}$-algebra.
\end{itemize}
We then have the following result, taken from the discussion in \cite{KhareWintenberger} before
Corollary~4.7.
\begin{proposition}
Assume $R$ is finite over $\mc{O}$ and each $R_v^{\Box}$ is Cohen--Macaulay. Then $R$ is flat over $\mc{O}$ and
Cohen--Macaulay. Furthermore, $R$ is Gorenstein if and only if each $R_v^{\Box}$ is.
\end{proposition}
\begin{proof}
By \cite[Prop.~4.1.5]{Kisin3}, we have a presentation
\begin{displaymath}
R^{\Box, \mathrm{univ}}=R^{\Box, \mathrm{univ}}_{\mathrm{loc}} \lbb x_1, \ldots, x_{r+n-1} \rbb/(f_1, \ldots, f_{r+s}),
\end{displaymath}
where $n=\# \Sigma$, $s=[F:\mathbf{Q}]$ and $r$ is a non-negative integer. Tensoring over $R^{\Box, \mathrm{univ}}_{\mathrm{loc}}$
with $R^{\Box}_{\mathrm{loc}}$ gives a presentation
\begin{displaymath}
R^{\Box}=R^{\Box}_{\mathrm{loc}} \lbb x_1, \ldots, x_{r+n-1} \rbb/(f_1, \ldots, f_{r+s}).
\end{displaymath}
This shows that $R^{\Box}$ has dimension at least $4n$. Since $R^{\Box}$ is a power series ring over $R$ in
$4n-1$ variables and $R$ is finite over $\mc{O}$, we see that $R$ has dimension 1 and $R^{\Box}$ has dimension $4n$.
Furthermore, writing $R^{\Box}=R\lbb T_1, \ldots, T_{4n-1} \rbb$, we see that $T_1, \ldots, T_{4n-1}, f_1, \ldots,
f_{r+s}, p$ is a system of parameters for $R^{\Box}_{\mathrm{loc}} \lbb x_1, \ldots, x_{r+n-1} \rbb$. Since each
$R^{\Box}_v$ is Cohen--Macaulay, so too is $R^{\Box}_{\mathrm{loc}} \lbb x_1, \ldots, x_{r+n-1} \rbb$, and it follows that
$T_1, \ldots, T_{4n-1}, f_1, \ldots, f_{r+s}, p$ is a regular sequence. This shows that $R^{\Box}$ is $\mc{O}$-flat
and Cohen--Macaulay, and furthermore that $R^{\Box}$ is Gorenstein if and only if each $R^{\Box}_v$ is. Finally, note
that $R^{\Box}$ is a power series ring over $R$, and so all these properties can be transferred to $R$.
\end{proof}
\begin{remark}
Finiteness of $R$ is known in many cases. When $\ol{\rho}$ is modular, one can often obtain finiteness of $R$ using
the Taylor--Wiles argument as modified by Kisin. Even without modularity it is often still possible to obtain
finiteness by using potential modularity.
\end{remark}
The following proposition is a very special case, showing how the above proposition can be combined with the
main results of this paper.
\begin{proposition}
Suppose $\Sigma$ consists exactly of the primes over $p$ and that for each $v \in \Sigma$ the local representation
$\ol{\rho} \vert_{G_{F_v}}$ is trivial. For $v \in \Sigma$, let $R^{\Box}_v$ be one of the rings $\wt{R}_{\ast}$
constructed in \S \ref{ss:rtilde}. Then, assuming $R$ is finite over $\mc{O}$, it is $\mc{O}$-flat, Cohen--Macaulay
and not Gorenstein.
\end{proposition}
\subsection{An $R=\mathbf{T}$ theorem}
Assume that $\ol{\rho}$ is modular. Then one can define a Hecke algebra $\mathbf{T}$ and a surjection $R \to \mathbf{T}$. The
original method of Taylor and Wiles shows that this map is an isomorphism in certain situations. Kisin's modification
of the method applies in greater generality, but only shows that $R[1/p] \to \mathbf{T}[1/p]$ is an isomorphism.
We simply remark here that if one knows that $R$ is $\mc{O}$-flat, then knowing that $R[1/p] \to \mathbf{T}[1/p]$ is an
isomorphism implies that the map $R \to \mathbf{T}$ is an isomorphism. Thus when $\ol{\rho}$ is trivial at the places
above $p$ and one uses the $\wt{R}_{\ast}$ local deformation conditions, one can expect to obtain an $R=\mathbf{T}$ theorem.
|
\section{The influence of the computer}
Even before computers were built, pioneers such as Babbage and Turing
realised that they would be designed on discrete principles, and would raise
theoretical issues which led to important mathematics.
Kurt G\"odel~\cite{godel}
showed that there are true statements about the natural numbers
which cannot be deduced from the axioms of a standard system
such as Peano's. This result was highly significant for the
foundations of mathematics, but G\"odel's unprovable statement
itself had no mathematical significance. The first example of
a natural mathematical statement which is unprovable in Peano
arithmetic was discovered by Paris and Harrington~\cite{ph}, and
is a theorem in combinatorics (it is a slight strengthening of
Ramsey's theorem). It is unprovable from the axioms because the
corresponding `Paris--Harrington function' grows faster than
any provably computable function. Several further examples of
this phenomenon have been discovered, mostly combinatorial in
nature.\footnote{Calculating precise values for Ramsey
numbers, or even close estimates, appears to be one of the
most fiendishly difficult open combinatorial problems.}
More recently, attention has turned from \emph{computability} to
\emph{computational complexity}: given that something can be
computed, what resources (time, memory, etc.) are required for
the computation. A class of problems is said to be
\emph{polynomial-time computable}, or in $\mathsf{P}$, if any
instance can be solved in a number of steps bounded by a polynomial
in the input size. A class is in $\mathsf{NP}$ if the same
assertion holds if we are allowed to make a number of lucky
guesses (or, what amounts to the same thing, if a proposed solution
can be checked in a polynomial number of steps). The great unsolved
problem of complexity theory asks:
\begin{center}
Is $\mathsf{P}=\mathsf{NP}$?
\end{center}
On 24 May 2000, the Clay Mathematical Institute announced a list of
seven unsolved problems, for each of which a prize of one million dollars
was offered. The $\mathsf{P}=\mathsf{NP}$ problem was the first on the
list~\cite{clay}.
This problem is particularly important for combinatorics since
many intractable combinatorial problems (including the existence
of a Hamiltonian cycle in a graph) are known to be in
$\mathsf{NP}$. In the unlikely event of an affirmative solution,
`fast' algorithms would exist for all these problems.
Now we turn to the practical use of computers.
Computer systems such as \textsf{GAP}~\cite{gap} have been developed,
which can treat algebraic or combinatorial
objects, such as a group or a graph, in a way similar to the
handling of complex numbers or matrices in more traditional
systems. These give the mathematician a very powerful tool for
exploring structures and testing (or even formulating) conjectures.
But what has caught the public eye is the use of computers to
prove theorems. This was dramatically the case in 1976 when
Kenneth Appel and Wolfgang Haken~\cite{ah} announced that they
had proved the Four-Colour Theorem by computer. Their announcement
started a wide discussion over whether a computer proof is really
a `proof' at all: see, for example, Swart~\cite{swart} and Tymoczko~\cite{tym}
for contemporary responses. An even more massive computation by Clement
Lam and his co-workers~\cite{lametal}, discussed by Lam in~\cite{lam},
showed the non-existence of a projective plane of order~$10$.
Other recent achievements include the classification of Steiner triple
systems of order~$19$~\cite{ko}.
Computers have been used in other parts of mathematics. For
example, in the Classification of Finite Simple Groups (discussed
below), many of the sporadic simple groups were constructed with
the help of computers. The very practical study of fluid dynamics
depends on massive computation. What distinguishes combinatorics?
Two factors seem important:
\begin{itemize}
\item[(a)] in a sense, the effort of the proof consists mainly in
detailed case analysis, or generate large amounts of data,
and so the computer does most of the work;
\item[(b)] the problem and solution are both discrete; the results
are not invalidated by rounding errors or chaotic behaviour.
\end{itemize}
Finally, the advent of computers has given rise to many new areas
of mathematics related to the processing and transmission of data.
Since computers are digital, these areas are naturally related
to combinatorics. They include coding theory (discussed below),
cryptography, integer programming, discrete optimisation, and
constraint satisfaction.
\section{The nature of the subject}
The last two centuries of mathematics have been dominated by the
trend towards axiomatisation. A structure which fails to satisfy the
axioms is not to be considered. (As one of my colleagues put it to
a student in a class, ``For a ring to pass the exam, it has to
get 100\%''.) Combinatorics has never fitted this pattern very well.
When Gian-Carlo Rota and various co-workers wrote an influential series of
papers with the title `On the foundations of combinatorial
theory' in the 1960s and 1970s (see~\cite{rota,cr}, for example),
one reviewer compared combinatorialists to nomads on the steppes
who had not managed to construct the cities in which other
mathematicians dwell, and expressed the hope that these papers
would at least found a thriving settlement.
While Rota's papers have been very influential, this view has not prevailed.
To see this, we turn to the more recent series on `Graph minors' by
Robertson and Seymour~\cite{rs}. These are devoted to the proof of a
single major theorem, that a minor-closed class of graphs is
determined by finitely many excluded minors. Along the way, a
rich tapestry is woven, which is descriptive (giving a topological
embedding of graphs) and algorithmic (showing that many graph
problems lie in $\mathsf{P}$) as well as deductive.
The work of Robertson and Seymour and its continuation is certainly
one of the major themes in graph theory at present, and has
contributed to a shorter proof of the Four-Colour Theorem, as well
as a proof of the Strong Perfect Graph Conjecture. Various authors,
notably Gerards, Geelen and Whittle, are extending it to classes of matroids
(see~\cite{ggw}).
What is clear, though, is that combinatorics will continue to
elude attempts at formal specification.
\section{Relations with mathematics}
In 1974, an Advanced Study Institute on Combinatorics was held at
Nijenrode, the Netherlands, organised by Marshall Hall and Jack van Lint.
This was one of the first presentations,
aimed at young researchers, of combinatorics as a mature mathematical
discipline. The subject was divided into five sections: theory of designs,
finite geometry, coding theory, graph theory, and combinatorial group theory.
It is very striking to look at the four papers in coding theory~\cite{hl}.
This was the youngest of the sections, having begun with the work
of Hamming and Golay in the late 1940s. Yet the methods being
used involved the most sophisticated mathematics: invariant theory,
harmonic analysis, Gauss sums, Diophantine equations.
This trend has continued. In the 1970s, the Russian school (notably Goppa,
Manin, and Vladut) developed links between coding theory and algebraic
geometry (specifically, divisors on algebraic curves). These links were
definitely `two-way', and both subjects benefited. More recently, codes
over rings and quantum codes have revitalised the subject and made new
connections with ring theory and group theory. In the related field of
cryptography, one of the most widely used ciphers is based on elliptic
curves.
Another example is provided by the most exciting development
in mathematics in the late 1980s, which grew
from the work of Vaughan Jones, for which he received a
Fields Medal in 1990. His research on traces of Von Neumann
algebras came together with representations of the Artin
braid group to yield a new invariant of knots, with
ramifications in mathematical physics and elsewhere. (See the
citation by Joan Birman~\cite{birman1} and her popular
account~\cite{birman2} for a map of this
territory.) Later, it was pointed out that the Jones polynomial is
a specialisation of the Tutte polynomial, which had been defined
for arbitrary graphs by Tutte and Whitney and generalised to matroids
by Tutte. Tutte himself has given two accounts of his discovery:
\cite{tutte1,tutte2}. The connections led to further work.
There was the work of Fran\c{c}ois Jaeger~\cite{jaeger}, who derived a
spin model, and hence an evaluation of the Kauffman polynomial,
from the strongly regular graph associated with the Higman--Sims
simple group; and that of Dominic Welsh and his collaborators
(described in his book~\cite{welsh}) on the computational complexity of
the new knot invariants.
Sokal~\cite{sokal} has pointed out that there are close relations between
the Tutte polynomial and the partition function for the Potts model in
statistical mechanics; this interaction has led to important advances in
both areas.
Examples such as this of unexpected connections, by their nature,
cannot be predicted. However, combinatorics is likely to be involved
in such discoveries: it seems that deep links in mathematics often
reveal themselves in combinatorial patterns.
One of the best examples concerns the ubiquity of the Coxeter--Dynkin
diagrams $A_n$, $D_n$, $E_6$, $E_7$, $E_8$. Arnol'd (see~\cite{arnold})
proposed finding an explanation of their ubiquity as a modern equivalent
of a Hilbert problem, to guide the development of mathematics.
He noted their occurrence in areas such as Lie algebras (the
simple Lie algebras over $\mathbb{C}$), Euclidean geometry (root systems),
group theory (Coxeter groups), representation theory (algebras of
finite representation type), and singularity theory (singularities
with definite intersection form), as well as their connection
with the regular polyhedra. To this list could be added
mathematical physics (instantons) and combinatorics (graphs
with least eigenvalue $-2$). Indeed, graph theory provides the
most striking specification of the diagrams: they are just the
connected graphs with all eigenvalues smaller than $2$.
Recently this subject has been revived with the discovery by Fomin and
Zelevinsky~\cite{fz} of the role of the ADE diagrams in the theory of cluster
algebras: this is a new topic with combinatorial foundations and applications
in Poisson geometry, integrable systems, representation theory and total
positivity.
Other developments include the relationship of
combinatorics to finite group theory. The Classification of Finite
Simple Groups~\cite{gorenstein} is the greatest collaborative effort
ever in mathematics, running to about 15000 journal pages. (Ironically,
although the theorem was announced in 1980, the proof contained a gap
which has only just been filled.)
Combinatorial ideas (graphs, designs, codes, geometries) were involved
in the proof: perhaps most notably, the classification of spherical
buildings by Jacques Tits~\cite{tits}. Also, the result has had a
great impact in combinatorics, with consequences both for symmetric
objects such as graphs and designs (see the survey by Praeger~\cite{praeger}),
and (more surprisingly) elsewhere as in Luks' proof~\cite{luks} that the graph
isomorphism problem for graphs of bounded valency is in $\mathsf{P}$.
This account would not be complete without a mention of the work of
Richard Borcherds~\cite{borcherds} on `monstrous moonshine', connecting
the Golay code, the Leech lattice, and the Monster simple group with
generalised Kac--Moody algebras and vertex operators in mathematical physics
and throwing up a number of product identities of the kind familiar from
the classic work of Jacobi and others.
\section{In science and in society}
Like any human endeavour, combinatorics has been affected by the great
changes in society last century. The first influence to be mentioned
is a single individual, Paul Erd\H{o}s, who is the subject of two
recent best-selling biographies~\cite{hoffman, schechter}.
Erd\H{o}s' mathematical interests were wide, but combinatorics was
central to them. He spent a large part of his life without a permanent
abode, travelling the world and collaborating with hundreds of mathematicians.
In the days before email, he was a vital communication link between
mathematicians in the East and West; he also inspired a vast body of research
(his 1500 papers dwarf the output of any other modern mathematician).
Jerry Grossman~\cite{grossman} has demonstrated the growth in multi-author
mathematical papers this century, and how Erd\H{o}s was ahead of this
trend (and almost certainly contributed to it).
Erd\H{o}s also stimulated mathematics by publicising his vast collection of
problems; for many of them, he offered financial rewards for solutions. As an
example, here is one of his most valuable problems. Let $A=\{a_1,a_2,\ldots\}$
be a set of positive integers with the property that the sum of the
reciprocals of the members of $A$ diverges. Is it true that $A$ contains
arbitrarily long arithmetic progressions? The motivating special case
(recently solved affirmatively by Green and Tao~\cite{gt}) is that where $A$
is the set of prime numbers: this is a problem in number theory, but
Erd\H{o}s' extension to an arbitrary set transforms it into combinatorics.
Increased collaboration among mathematicians goes beyond the influence of
Erd\H{o}s; combinatorics seems to lead the trend. Aspects of this trend
include large international conferences (the Southeastern Conference on
Combinatorics, Graph Theory and Computing, which held its 42nd meeting in 2011,
attracts over 500 people annually), and electronic journals (the
\textit{Electronic Journal of Combinatorics}~\cite{ejc}, founded in 1994,
was one of the first refereed specialist electronic journals in mathematics).
Electronic publishing is particularly attractive to combinatorialists.
Often, arguments require long case analysis, which editors of traditional
print journals may be reluctant to include in full.
On a popular level, the Sudoku puzzle (a variant of the problem of completing
a critical set in a Latin square) engages many people in combinatorial
reasoning every day. Mathematicians have not been immune to its attractions.
At the time of writing, MathSciNet lists 38 publications with `Sudoku' in
the title, linking it to topics as diverse as spreads and reguli, neural
networks, fractals, and Shannon entropy.
Our time has seen a change in the scientific
viewpoint from the continuous to the discrete. Two mathematical
developments of the twentieth century (catastrophe theory and
chaos theory) have shown how discrete effects can be produced by
continuous causes. (Perhaps their dramatic names reflect the
intellectual shock of this discovery.) But the trend is even
more widespread.
In their book introducing a new branch of discrete mathematics
(game theory), John von Neumann and Oskar Morgenstern~\cite{nm}
wrote:
\begin{quote}
The emphasis on mathematical methods seems to be shifted more
towards combinatorics and set theory -- and away from the
algorithm of differential equations which dominates mathematical
physics.
\end{quote}
How does discreteness arise in nature? Segerstr{\aa}le~\cite{us}
quotes John Maynard-Smith as saying
``today we really
do have a mathematics for thinking about complex
systems and things which undergo transformations
from quantity into quality''
or from continuous to discrete, mentioning Hopf bifurcations as
a mechanism for this.
On the importance of discreteness in nature, Steven Pinker~\cite{pinker}
has no doubt. He wrote:
\begin{quote}
It may not be a coincidence that the two systems in the universe
that most impress us with their open-ended complex design --
life and mind -- are based on discrete combinatorial systems.
\end{quote}
Here, `mind' refers primarily to language, whose combinatorial
structure is well described in Pinker's book. `Life' refers to
the genetic code, where DNA molecules can be regarded as words
in an alphabet of four letters (the bases adenine, cytosine,
guanine and thymine), and three-letter subwords encode amino acids,
the building blocks of proteins.
The Human Genome Project, whose completion was announced in 2001,
was a major scientific enterprise to
describe completely the genetic code of humans.
(See~\cite{bbklp} for an account of the mathematics involved, and
\cite{lander2} for subsequent developments.)
At Pinker's university (the Massachusetts
Institute of Technology), the Whitehead Laboratory was
engaged in this project. Its director, Eric Lander, rounds off
this chapter and illustrates its themes. His doctoral thesis~\cite{lander}
was in combinatorics, involving a `modern' subject (coding theory),
links within combinatorics (codes and designs), and links to other
parts of mathematics (lattices and local fields). Furthermore,
he is a fourth-generation academic descendant of Henry Whitehead.
But there are now hints that discreteness plays an even more fundamental
role. One of the goals of physics at present is the construction of a theory
which could reconcile the two pillars of twentieth-century physics, general
relativity and quantum mechanics. In describing string theory,
loop quantum gravity, and a variety of other approaches including
non-commutative geometry and causal set theory, Smolin~\cite{smolin} argues
that all of them involve discreteness at a fundamental level (roughly the
Planck scale, which is much too small and fleeting to be directly observed).
Indeed, developments such as the holographic principle suggest that the basic
currency of the universe may not be space and time, but information,
measured in bits. Maybe the `theory of everything' will be combinatorial!
|
\section{Introduction}\label{intro}
Topological encoding of quantum data enables computation to be protected from the effects of decoherence on qubits and of physical device errors in processing. A logical qubit is encoded in the entangled state of many physical qubits; the exact ratio is determined by the \emph{code distance}, which is chosen based on measured physical error rates and desired logical error rates. As long as physical errors are below the \emph{threshold value}, increasing the number of physical qubits can exponentially suppress error on the logical qubits \cite{threshold}. Of the many types of codes known, the \emph{surface code} stands out as having the highest tolerance of component error ($\sim 1\%$ in recent results) when implemented on a simple 2-dimensional lattice of qubits with
nearest-neighbour interactions \cite{topo-q-memory,raussendorfprl,austin1,austin14,new-threshold-austin,bombin,bomb11}.
Surface codes were first introduced by Kitaev \cite{kitaev} in the context of anyonic quantum computing. There are two implementation strategies for such codes: the first uses exotic anyonic particles \cite{anyonrev}, and the second takes an active approach to error correction on a lattice of regular qubits. It is the latter that we are concerned with here. Within the active implementation, the first type of surface code that was developed was the \emph{planar code}: each logical qubit occupies a separate
code surface, with its own boundaries \cite{planar-bk,planar-fm}. While the error correction requires only nearest-neighbouring (NN) physical qubits to interact, multi-qubit gate operations were proposed to be performed transversally between surfaces. By contrast, the now standard surface code defines logical qubits as \emph{defects} (introduced degrees of freedom) within a single lattice, and deformation and braiding of the defects within a single surface performs gates between them. The price of maintaining NN interactions is over 3
times the number of physical qubits per logical qubit \cite{raussendorfprl,austin1}.
The reduced qubit requirements of the planar code make it very attractive for nearer-term experimental implementations of the surface code; for example the smallest correctable code (distance 3) uses 13 physical qubits for a single planar logical qubit, but a standard (double-defect) surface code uses 72. More broadly, smaller resource requirements are extremely useful for applications where small numbers of logical qubits need to be communicated in order to take part in distributed computing. However, the requirement for transversal two-qubit gates has previously made a planar encoding unfeasible for many systems where the physical qubits are confined in 2D and subject only to NN interactions, such as quantum dots \cite{qdos-architecture,optics-dots}, superconducting qubits \cite{super-archi,super-archi-austin}, trapped atoms \cite{atomic-archi}, nitrogen-vacancy (NV) diamond arrays \cite{2d-ss-architecture}, and some ion trap architectures \cite{2d-rf-iontraps,2d-penning-iontraps}. Until now, the only way to maintain NN interactions when performing multiple logical qubit gates was to move to a
defect-based surface code scheme.
In this paper we solve this problem by introducing a new method of deforming and combining planar code surfaces which we term \emph{lattice surgery}. By analogy with the term used in geometric topology, lattice surgery comprises the ``cutting" and ``stitching" of code surfaces to produce other planar surfaces. We show that these operations on the planar code produce novel code operations while requiring only standard NN physical interactions, and maintaining full fault-tolerance. Furthermore we demonstrate that these new operations can be combined to produce multiple logical qubit gates without any transversal interactions, and we give the full construction for a \leavevmode\hbox{\footnotesize{CNOT }} operation between two planar qubits. To complete the universal gate set we demonstrate how magic states can be injected into the code space, and also give a useful direct construction of the Hadamard gate. We show how defect- and planar- based qubits can be interchanged, and demonstrate how a defect-based qubit can be detached from a code surface as a planar qubit. We finish by detailing two important medium-term achievable experiments that could be performed using lattice surgery: producing entangled planar qubits, both as Bell pairs and GHZ states; and a full NN \leavevmode\hbox{\footnotesize{CNOT }} between two distance-3 planar qubits. These use significantly fewer physical resources than defect-based codes, with our smallest lattice surgery \leavevmode\hbox{\footnotesize{CNOT }} requiring 53 qubits to implement: half the physical qubits of the smallest known defect-encoded \leavevmode\hbox{\footnotesize{CNOT }} operation, which we also describe.
\section{Surface codes}\label{scs}
\begin{figure}[t]
\centering
\includegraphics[width=9cm]{lattice-and-stabs-syndrome.jpg}
\caption{Part of the basic lattice of the surface code. Data qubits are shown large, syndrome qubits as small. The label `A' marks a face plaquette, `B' a vertex plaquette.}\label{plaquettes}
\end{figure}
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=6cm]{ZCN}
\label{zcn}}\hspace{1cm}
\subfigure[]
{\includegraphics[width=8cm]{XCN}
\label{xcn}}\hspace{1cm}
\caption{Circuits for syndrome extraction: a) Z-syndrome b) X-syndrome. Measurements are in the computational basis.}
\label{syndromes}
\end{figure}
Surface codes use a two-dimensional regular lattice of entangled physical qubits to give the substrate on which logical qubits are defined \cite{planar-bk,planar-fm}. The lattice is made up of two types of qubits, data and syndrome, differing only in their function within the code. Syndrome qubits are
repeatedly and frequently interacted with neighbouring data qubits and
measured to detect the presence of errors. Data qubits are measured
less frequently and only to perform computation. The qubits are arranged in a lattice as in figure \ref{plaquettes}. The lines in the figure are aids to the eye, and do not designate any physical interactions or structures. The data qubits are in a simultaneous eigenstate of Pauli-$Z$ operators around each face (for example the ``plaquette" $A$ in figure \ref{plaquettes}), and Pauli-$X$ around each vertex (e.g.. $B$ in figure \ref{plaquettes}). That is, the \emph{stabilizers} of the systems are, for all face $F$ and vertex $V$ plaquettes,
\begin{equation} \otimes_{i \in F} Z_i \ \ \ \ \mathrm{and} \ \ \ \ \ \otimes_{j \in V} X_j\label{lattstate}\end{equation}
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=5cm]{planar-qubit-syndrome.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=5cm]{single-defect-qubit.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=8cm]{double-defect-qubit.jpg}
}\hspace{1cm}
\caption{The three surface code methods of encoding a single logical qubit, shown here at distance 4: a) planar; b) single defect; c) double defect. Sample logical operators are marked in each case.}
\label{distfour}
\end{figure}
\noindent where $\{i \in F\}$ and $\{j \in V\}$ are the data qubits in each face and vertex plaquette respectively.
The job of the syndrome qubits is to keep the lattice in a simultaneous eigenstate of these operators, in the face of possible qubit and measurement errors. To do this, syndrome qubits are placed in the centre of each plaquette, and in each round of error correction measure the 4-party stabilizer of that plaquette. The standard circuits for these \emph{syndrome measurements} are given in figure \ref{syndromes}. In the absence of the errors, the syndrome measurement will have
the same value in every round. Every time the syndrome measurement
changes value, an endpoint of a local chain of errors has been
detected. A pattern of chains of corrective operations highly likely
to maintain the correct logical state can be inferred using
Edmonds' minimum weight perfect matching algorithm \cite{Edmo65a,Edmo65b,Kolm09}.
Corrective operations are applied to classical data associated with
the measurement results and algorithm being executed, rather than the physical qubits. This ensures
that no corrective operations need to be applied to the qubits,
reducing the quantum error rate.
Qubits are defined using spare degrees of freedom in this lattice. In general, the requirement of the stabilizers (\ref{lattstate}) fixes the state fully; there are, however, several ways of introducing a degree of freedom into the lattice state. The first method is to define the lattice as having both \emph{rough} and \emph{smooth} boundaries as in figure \ref{distfour}(a). This introduces a single degree of freedom, and so the entire lattice can be used to encode a single logical qubit. This is the planar version of the surface code. A second method produces the required degree of freedom by not enforcing one of the stabilizers of equation (\ref{lattstate}) -- that is, introducing a \emph{defect} into the lattice, figure \ref{distfour}(b). In both cases we can define the logical operators of the encoded qubit, shown in the figures. In practice, in the defect-based code the qubits are defined by double defects, figure \ref{distfour}(c), as this localises the logical operators (no error chains can go between the defects and the edge of the lattice as they are opposite-type boundaries) and allows many qubits to be easily defined on a single surface. A third method involves introducing ``twists" to the lattice \cite{bombtwist}.
\begin{figure}[t]
\centering
\includegraphics[width=8cm]{transversal.jpg}
\caption{Transversal logical \leavevmode\hbox{\footnotesize{CNOT }} operation between two planar logical qubits. The pink interactions denote \leavevmode\hbox{\footnotesize{CNOT }} operations between pairs of physical qubits. Syndrome qubits have been suppressed for clarity.}\label{transversal}
\end{figure}
In both planar and defect cases, a logical error is a chain of single-qubit errors that mimics a logical operator. Such error chains are undetectable, and cause a failure of the code. The \emph{code distance} is the measure of strength of the code, and is the length of the smallest undetectable error chain -- that is, the length of the smallest logical operator. For example, a code of distance 3 has the smallest logical operator as three physical qubit operations. Such a surface code can detect and correct a single physical error. Owing to their construction, a planar qubit of distance $d$ is generally around three times smaller than a defect-based qubit of the same distance. For both code types, $d$ rounds of error correction are needed fully to correct the lattice, in order to produce a spatio-temporal cell of depth $d$ to perform minimum weight matching of the ends of error chains \cite{raussendorf3D}. Whenever operations are performed on an encoded qubit, these $d$ rounds are required to correct the lattice before moving on.
Surfaces can initially be prepared in either the logical $\ket{0}$ or logical $\ket{+}$ state. To prepare a $\ket{0}_l$, all physical qubits are prepared in the $\ket{0}$ state, and then $d$ rounds of syndrome measurements performed to ensure fault-tolerance. Similarly, a $\ket{+}_L$ state is created by preparing all qubits in the physical $\ket{+}$ state and then performing syndrome measurements. At the end of the computation, the value of a logical qubit is read out by measuring all the qubits comprising the logical qubit in the measurement basis ($X$ or $Z$). These measurements are then subject to error correction, and the result of the logical measurement read from the parity of the logical operators measured.
The standard methods for performing two-qubit gate operations (usually the \leavevmode\hbox{\footnotesize{CNOT }} operation) differ significantly between planar and defect-based surface codes. For a planar \leavevmode\hbox{\footnotesize{CNOT }} the original method was transversal: logical qubits are defined by construction on separate surfaces, and each physical qubit of one surface performs a \leavevmode\hbox{\footnotesize{CNOT }} with the corresponding physical qubit of the other surface, as shown in figure \ref{transversal}. After these operations, a \leavevmode\hbox{\footnotesize{CNOT }} has occurred between the logical qubits. By contrast, in the defect-based code there are no transversal operations, and the \leavevmode\hbox{\footnotesize{CNOT }} is performed by \emph{braiding} defects: extending a defect by measuring out qubits in a line, and passing this extended defect around the second logical qubit defect \cite{raussendorfprl}. The only operations other than measuring out individual qubits are those of the standard error correction procedure. A second method of performing such a gate is \emph{code deformation}: the boundaries of the code lattice itself are deformed around the defects, performing interactions on the logically encoded qubit \cite{bombin,bomb06,bomb11}.
The use of NN-only interactions for full computation has made the defect-based code the method of choice, as transversal operations create many more implementation problems than NN interactions. The new procedure of \emph{lattice surgery} that we introduce in this paper removes transversal two-qubit operations from the planar code, and not only allows a NN only \leavevmode\hbox{\footnotesize{CNOT }} operation to be performed, but also introduces new logical qubit operations (which we term ``split" and ``merge") in a manner that is practical for a system that may ultimately be built
\section{Lattice surgery}\label{lsurgery}
The standard methods for implementing surface code gates non-transversally treat the lattice in ways familiar from algebraic topology \cite{algetop}, continuously deforming the lattice in order to achieve the required results. The methods proposed here break from this by introducing discontinuous deformations of the lattice, analogous to the operations of surface surgery in geometric topology (see for example \cite{mono}). We introduce the notions here of ``merging" and ``splitting" planar code lattices, and demonstrate the logical operations that they perform on the encoded data.
Lattice merging occurs when two code surfaces become a single surface. This is implemented by measuring joint stabilizers across the boundaries of the surfaces during error correction cycles. Depending on which boundaries are joined, this operation behaves differently. Splitting of code surface is the opposite procedure, in which joint stabilizers are cut, forming extra boundaries that turn one code surface into two. Again, the types of boundaries created determine the exact nature of the final states. We will now demonstrate in detail the results of these four operations.
\subsection{Lattice merging}\label{lmerge}
Let us consider the system shown in figure \ref{merge1}. There are two logical planar code surfaces, each separately stabilized and encoding a single logical qubit, and a row of `intermediate' uninitialised physical qubits. We merge the two systems by first preparing the intermediate data qubits in the state $\ket{0}$, and then performing $d$ rounds of error correction, treating the entire system as a single data surface.
\begin{figure}[t]
\centering
\includegraphics[width=10cm]{merge-stabilizers.jpg}
\caption{Arrangements of physical qubits for rough lattice merging. Left and right continuous surfaces encode separate logical qubits. The pink qubits form the intermediate qubit line for the merging operation.}\label{merge1}
\end{figure}
After correction, we will be able to reliably determine the sign of
the new $X$-stabilizer measurements spanning the old boundary. Note that
only the stabilizer measurements performed in the first round of the d
rounds will be known reliably, since only these are buried under
sufficient additional information to enable reliable correction. Later
rounds of stabilizer measurements become reliably known only as still
further rounds of error correction are performed.
Armed with reliable $X$-stabilizer measurements spanning the boundary,
we can reliably infer the eigenvalue of the product of these
$X$-stabilizers, which is equivalent to the two-logical-qubit operator
$X_LX_L$. The merge procedure described above is thus equivalent to
measuring $X_LX_L$. We define this merging operation as a rough merge:
the rough boundaries of the two surfaces are merged together. This
leads us to define a second type, that of a smooth merge, where it is
the smooth boundaries that are the subject of the merge operation. In
the case of a smooth merge, the intermediate qubits are prepared in
the $\ket{+}$ state before the new joint operators are measured. By
symmetry, a smooth merge is equivalent to measuring $Z_LZ_L$.
The lattice that remains now potentially has a sequence of syndrome measurements down the join that are incorrect. If there are an even number, then they can be corrected in the usual way by joining pairs with chains of $Z$ operations, as in \cite[\S V]{austin1}. In the case of an odd number of incorrect syndromes, the first one is not corrected, but simply tracked in software through the calculation (so that all subsequent correction operations correct to this ``incorrect" value). In this case we choose a particular $Z_L$ logical operator chain to be our ``reference" chain for that qubit; as long as the position of the chain is stored in memory and used for subsequent calculation and measurement, the logical qubit remains in the correct state.
This action of measuring $X_LX_L$ on the two original qubits makes two things happen. Firstly, the state after measurement is non-deterministic, and correlated to the measurement outcome. Secondly, the planar surface now only has a single qubit degree of freedom, so we require a mapping from the original logical qubit states to the new logical qubit state post-merge. Let us consider the case where $\ket{\psi} = \alpha \ket{0}_L + \beta\ket{1}_L$ is merged with $\ket{\phi} = \alpha^\prime \ket{0}_L + \beta^\prime \ket{1}_L$. The outcome of merging is the measurement of $X_LX_L$; were this performed on two separate qubits then the state after the measurement is
\begin{equation} \frac{1}{\sqrt{2}}\left( \ket{\psi}\ket{\phi} + (-1)^M \ket{\bar{\psi}}\ket{\bar{\phi}} \right) \label{mergstate}\end{equation}
\noindent where $\ket{\bar{A}} = \sigma_x\ket{A}$, and $M$ is the outcome of the logical measurement, 0 or 1.
As is usual in surface code work, we now correct for the outcome of the measurement, leaving us with a deterministic state after the correction is applied. The correction will in practice be ``applied" by changing the interpretation of subsequent measurement outcomes, rather than by physical state operations. To see which corrections need applying, let us expand and re-write equation (\ref{mergstate}) dependent on the measurement outcome:
\begin{eqnarray} (\alpha\alpha^\prime + \beta\beta^\prime)(\ket{00}_L + \ket{11}_L) + (\alpha\beta^\prime + \beta\alpha^\prime)(\ket{01}_L + \ket{10}_L) & {\ \ \ \ \ \ } & M=0\nonumber \\
(\alpha\alpha^\prime - \beta\beta^\prime)(\ket{00}_L - \ket{11}_L) + (\alpha\beta^\prime - \beta\alpha^\prime)(\ket{01}_L - \ket{10}_L) & {\ \ \ \ \ \ } & M=1\label{ems}\end{eqnarray}
We also now need to consider the mapping to the new post-merge single qubit. A logical $\ket{0}_L$ state of the new surface will be the even-parity state of the combined $Z_L$ operator -- which is the simple product of the original $Z_L$ operators. The logical $\ket{1}_L$ will be the odd-parity state of the product of the original $Z_L$ operators. However, we can see from equation (\ref{ems}) that the combinations of odd and even parity states that we have differ depending on the measurement result of the merge. We therefore have a conditional mapping, based on the $X_LX_L$ measurement outcome:
\begin{eqnarray}\ket{0}_L & \longrightarrow & \frac{1}{\sqrt{2}}(\ket{00}_L + (-1)^M \ket{11}_L) \nonumber \\
\ket{1}_L & \longrightarrow & \frac{1}{\sqrt{2}}(\ket{01}_L + (-1)^M \ket{10}_L)\label{mnb} \end{eqnarray}
If we now write the merge operation using the symbol ``$\leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ }$", using equations (\ref{ems}) and this mapping, we find
\begin{eqnarray} \ket{\psi} \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } \ket{\phi} & = & \alpha \ket{\phi} + (-1)^M \beta \ket{\bar{\phi}} \nonumber \\
{} & = {} & \alpha^\prime \ket{\psi} + (-1)^M \beta^\prime \ket{\bar{\psi}}\label{stabrep}\end{eqnarray}
In classical terms, the truth table for the merge operation between qubits either in the state $\ket{0}_L$ or the state $\ket{1}_L$ is an \leavevmode\hbox{\footnotesize{XOR }}:
\begin{equation}
\centering
\begin{array}{c c | c}
\mathrm{In(1)} & \mathrm{In(2)} & \mathrm{Out}\\
\hline
0 & 0 & 0\\
0 & 1 & 1\\
1 & 0 & 1\\
1 & 1 & 0\end{array}
\end{equation}
In the case of a smooth merge, the intermediate qubits are prepared in the $\ket{+}_L$ state before the new joint operators are measured. It is then the logical-$X$ operators that come merged, and so the action of the merge is an \leavevmode\hbox{\footnotesize{XOR }} in the Hadamard basis of the qubits:
\begin{eqnarray} \ket{\psi} \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } \ket{\phi} & = & (a \ket{+}_L + b\ket{-}_L) \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } ( a^\prime \ket{+}_L + b^\prime\ket{-}_L)\nonumber\\
{} & = & a^\prime\ket{\psi} + (-1)^Mb^\prime \ket{\bar{\psi}}\nonumber\\
{} & = & a\ket{\phi} + (-1)^Mb\ket{\bar{\phi}}
\end{eqnarray}
\ref{merge} illustrates the explicit transformations for two distance 2 codes. In the case of both smooth and rough merges, in order to preserve full fault-tolerance of the surface (and to gain a correct value for the logical $X_LX_L$ or $Z_LZ_L$ measurement), $d$ rounds of error correction are needed for a distance $d$ code to give the correct spatio-temporal volume for minimum weight matching of errors. Note that the distance of the merged surface in this configuration is the same as of the original surfaces: the length of the smallest error chain remains the same. Only if the merge increases the smallest error chain length does the code distance increase. This merge operation that we have defined has one very interesting property that sets it apart from other operations used on the surface code. While it is well-defined, fault-tolerant, and preserves the code space, it is not a unitary operation in the logical space. Two logical qubits are input into the operation, but only one emerges.
\subsection{Lattice splitting}\label{lsplit}
The second type of new code operation, \emph{lattice splitting} is, in a sense, the converse operation to merging. A single logical qubit surface is split in half by a row of measurements that remove data qubits from the lattice. This leaves two separately stabilised surfaces at the end of the operation. As with merging there are two types, depending on the boundary along which the split occurs.
\begin{figure}[t]
\centering
\includegraphics[width=10cm]{smooth-split.jpg}
\caption{Arrangements of qubits for smooth lattice splitting. The pink qubits are measured out in the $X$-basis to leave two separately-stabilized logical qubit surfaces.}\label{split1}
\end{figure}
Let us consider first the \emph{smooth split}, shown in figure \ref{split1}. The middle row of qubits, shown, is measured out in the Pauli-$X$ basis. This is the same configuration as for a smooth merge, in which case the marked qubits would be the intermediate qubits, initialised in the $\ket{+}$ state. As with the merge, we can see the action of the split operation through the effect on the
logical operators.
In the case of the smooth split, after measuring out the intermediate qubits we are left with two surfaces that are then individually stabilized, as before for a total of $d$ rounds of error correction, where $d$ is the code distance. Unlike in the merge case, splitting can change the code distance: a split that divides a square surface symmetrically will halve the code distance. To end up with two surfaces of distance $d$, then, we need to start with one surface of size $d\times 2d$ (which also has code distance $d$). After the split has been performed, we can look at the new plaquette operators on the join. Firstly, we can see that none of the joint $Z$ operators change at all: measuring out the qubits removes a row of face plaquettes from the error correction entirely, and leaves the surrounding face plaquettes untouched. The states of all three qubits before and after the split are therefore in the same superposition of eigenstates of the $Z$ logical operator.
The action of the split on the $X$ logical operator is more complicated. The set of $X$-measurements on the row of qubits will each have a random outcome, 0 or 1. Each measurement leaves 3-qubit $XXX$ plaquettes on either side of split, and the parity that then needs to be tracked for the purposes of error correction is the product of the measurement of this stabilizer with the measurement outcome at the split of what was the 4th qubit in the plaquette. The parity of the logical $X$ operator of the state of the remainder of the surface nevertheless remains fixed. However, as we now have two separate surfaces, this logical state is distributed across the two surfaces: the two surfaces are defined by reference to a single joint logical operator, rather than their individual ones. That is, the two surfaces will be in an entangled state of their logical $X$ operators.
It is important to note that the action of the split means that, for the individual surfaces, there is no longer full equivalence between each set of $X_i$ operators across the surface that can make up an $X_L$ operator. This can be dealt with in two ways. First, a single line of $X_i$ operators is defined to be the $X_L$ operator (for example, individual $X$ operators on the top row of qubits). As long as this definition is the same on the two surfaces, subsequent computational use of these qubits will preserve their state. The second alternative is to correct the surface based on the measurement results so that all putative $X_L$ operators again become equivalent. This may be done by ``pairing" those 3-term stabilizers on the split that have negative parity, as in the merging. This may be performed either by physically performing the pairing operations (in practise not an efficient method as additional errors may be introduced), or by keeping track of these operations in the interpretation of future measurement results.
We can therefore write out the action of a smooth split, which preserves the logical $Z$ operator but splits the logical $X$ over the two qubits produced, once the necessary corrections or logical operator definitions have been arranged:
\begin{equation} \alpha\ket{0}_L + \beta \ket{1}_L \longrightarrow \alpha\ket{00}_L + \beta \ket{11}_L\label{stabsplit}\end{equation}
By exchanging $X$ and $Z$ in the above argument, we can find the result of a rough split, which preserves the logical $X$ operator but splits the logical $Z$:
\begin{equation} a\ket{+} + b \ket{-} \longrightarrow a\ket{++} + b \ket{--}\end{equation}
As with the merge operation, the split operation is not unitary as the number of qubits has not been preserved. However, unlike the case of a merge, information has not been lost at the logical level: the original state on a single qubit can be recovered logically by performing a reversing merge operation after the split. Distributing a state across two logical qubits as the split does is indeed a common feature of surface codes. This is exactly what is done when a double defect is used to encode a single logical qubit: in fact what is encoded is an entangled pair, which is itself an encoding of a single qubit. \ref{split} illustrates the explicit transformations for two distance 2 codes.
\section{Universal gate operations with lattice surgery}\label{universal}
We have now defined the operations of lattice surgery, splitting and merging the code surfaces. What is not immediately clear from the definitions is whether these operations are universal for quantum computing in the logical space of the planar surface code. We now demonstrate that this is in fact the case, by constructing a standard universal set comprising a logical \leavevmode\hbox{\footnotesize{CNOT }} gate and arbitrary logical single-qubit rotations (using magic state distillation and injection, as in the case of the standard surface code). We also give a method for performing the Hadamard gate as a basic code operation.
\subsection{The \leavevmode\hbox{\footnotesize{CNOT }} gate}\label{cnot}
\begin{figure}[t]
\centering
\includegraphics[width=9cm]{cnot-1-stabilizers.jpg}
\caption{Layout of qubits for a \leavevmode\hbox{\footnotesize{CNOT }} operation with lattice surgery. Control (C) and target (T) surfaces interact by merging and splitting with the intermediate surface (INT).}\label{unicnot}
\end{figure}
The construction of a full \leavevmode\hbox{\footnotesize{CNOT }} gate using lattice surgery is shown in figure \ref{unicnot}. We start with the two logical qubits of distance $d$ that are the control, in state
\begin{equation} \ket{C} =\alpha\ket{0} + \beta\ket{1} = \bar{\alpha}\ket{+}_L + \bar{\beta}\ket{-}_L \end{equation}
\noindent and the target, in state
\begin{equation} \ket{T} = \alpha^\prime\ket{0}_L + \beta^\prime\ket{1}_L \end{equation}
\noindent We also have an intermediate logical qubit surface of distance $d$ initialised to the logical $\ket{INT} = \ket{+}_L$ state, and two strips of physical qubits, the first all in the physical $\ket{+}$ state and the second all in the physical $\ket{0}$ state, to help with merge operations.
The first step is to smooth merge the surfaces $C$ and $INT$. After the conditional definition of logical state post-merge, this creates a single surface in the state
\begin{eqnarray} \ket{C} \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } \ket{INT} & = & \bar{\alpha}\ket{+}_L + (-1)^M \bar{\beta}\ket{+}_L \nonumber\\
{} & = & \frac{1}{\sqrt{2}}\Big( \bar{\alpha} + (-1)^M \bar{\beta}\Big) \ket{0} + \frac{1}{\sqrt{2}}\Big( \bar{\alpha} - (-1)^M \bar{\beta}\Big) \ket{1}
\end{eqnarray}
\noindent where, as before, $M$ is the measurement outcome of the $Z_LZ_L$ operator performed during the merge. If $M$ is even then we redefine the basis of this qubit by a bit-flip (this need not be performed, but simply tracked in software). We then have
\begin{equation} \ket{C} \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } \ket{INT} = \alpha\ket{0} + \beta \ket{1}\end{equation}
For full fault-tolerance, $d$ rounds of error correction are performed to create this merge operation. The code distance after the merge is still $d$. We then split this new surface back into the original two surfaces with a smooth split, measuring out the qubits along the split in the $X$ basis. With the logical-$X_L$ operators either redefined or the surface corrected, the state of these two new surfaces is then
\begin{equation} \ket{C^\prime \ INT^\prime} = \alpha\ket{00}_L + \beta\ket{11}_L\end{equation}
\noindent The code distance of both qubits is again $d$, and $d$ rounds of error correction are required for this step as well. We now perform a rough merge operation between the surfaces $INT$ and $T$:
\begin{eqnarray} \ket{C^\prime \ (INT^\prime \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } T) }& = & \alpha\ket{0}_L\otimes ( \ket{0}_L \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } \ket{T} ) + \beta \ket{1}_L \otimes(\ket{1}_L \leavevmode\hbox{\kern3.1pt\tiny{M}\kern-8pt\footnotesize$\bigcirc$ } \ket{T})\nonumber\\
{} & = &{} \alpha\ket{0}_L\otimes \ket{T} + (-1)^{(M^\prime)} \beta \ket{1}_L \otimes \ket{\bar{T}}\end{eqnarray}
\noindent where $M^\prime$ is again the merge measurement outcome, and again we have redefined the logical states dependent on the merge measurement outcome. This operation also takes $d$ rounds of error correction, and at the end we are left with two logical qubits of distance $d$. The outcome of this operation, as can be seen from the above equation, is a fully reversible \leavevmode\hbox{\footnotesize{CNOT }} operation. Fault tolerance has been maintained throughout with the multiple rounds of error correction, and the fact that at no point during this operation has the code distance of any of the logical surfaces used dropped below $d$. We have therefore constructed a fault-tolerant, unitary \leavevmode\hbox{\footnotesize{CNOT }} operation from lattice surgery operations without transversal gates.
\subsection{State injection}\label{stateinjection}
The second element that we need for a universal gate set is state injection. This allows for both state preparation and also arbitrary single-qubit rotations, in accordance with standard techniques \cite{magic-dist}. As with all surface code implementations, rotations on the code surface cannot in general be performed by manipulating only the code surface of the single qubit. Arbitrary rotations require the use of ancilla states, with \leavevmode\hbox{\footnotesize{CNOT }} operations between the ancilla and logical qubit in order to implement a rotation gate on the logical qubit.
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=4cm]{injecta.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=4cm]{injectb.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=4cm]{injectc.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=4cm]{injectd.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=4cm]{injecte.jpg}
}
\caption{Injecting an arbitrary state $\ket{\Psi} = \alpha \ket{0} + \beta \ket{1}$ into a planar surface: a) the blue qubits are prepared in \ket{0}, and the pink is the entangled state $\ket{\Psi}$; b) \leavevmode\hbox{\footnotesize{CNOT }} operations are performed to create $\alpha \ket{000} + \beta\ket{111}$ on the pink qubits; c) measuring stabilizers gives a distance-3 surface in the logical $\ket{\Psi}$ state; d) prepare the blue qubits in \ket{0}; e) merging in the blue qubits gives a distance-4 surface in $\ket{\Psi}$.}
\label{inject}
\end{figure}
As we now have a \leavevmode\hbox{\footnotesize{CNOT }} operation, a standard procedure can be used to perform these gates on the planar code, given a supply of suitable ancilla states. These ancilla states need to be topologically protected, so we require a state injection procedure for magic states. The general procedure was given for the planar code in \cite{topo-q-memory}, and a worked example for general surface codes in \cite[\S VI(C)]{austin1}. We here give the exact operations that would be used in a lattice surgery context.
Figure \ref{inject} demonstrates the procedure for injecting $ \alpha \ket{0} + \beta \ket{1}$. First a distance-3 logical qubit surface is prepared, with all qubits except one data qubit in the \ket{0} state. The remaining qubit is the magic state to be injected. \leavevmode\hbox{\footnotesize{CNOT }} operations are performed between this qubit and the syndrome qubits immediately above and below it. These syndrome qubits are then swapped with the data qubits immediately above or below to create the 3-qubit state $\alpha \ket{000} + \beta\ket{111}$, figure \ref{inject}(b). This surface is then stabilized, to give a distance-3 logical qubit in the logical state $ \alpha \ket{0}_L + \beta \ket{1}_L$, figure \ref{inject}(c). Additional physical qubits are prepared in \ket{0}, figure \ref{inject}(d), and then merged into the original logical qubit to give a distance-4 logical qubit in state $ \alpha \ket{0}_L + \beta \ket{1}_L$, figure \ref{inject}(e). The surface may be increased to any desired code distance by further merging.
\subsection{The Hadamard gate}
A useful element in the defect-based surface code is the ability to perform Hadamard gates without needing an Euler decomposition \cite[\S VII]{austin1}. The procedure for creating a Hadamard gate in the planar code differs as there is not a fixed background lattice, and we now give a method for performing this operation.
Performing a Hadamard gate transversally by Hadamard operations on each individual qubit will leave us with a planar qubit that is in the correct state (an eigenstate of $X_L$ is taken to the corresponding eigenstate of $Z_L$ and \emph{vice versa}), but it will leave the planar surface at a different orientation from the original, figure \ref{had1}. If there is no further processing to be done on the qubit then this can remain. Alternatively, if the physical interconnects between planar surfaces are movable, then they can be rotated through 90 degrees on all subsequent interactions. However, if the underlying physical implementation has fixed qubits (or we wish them to remain so for reasons of scalability), then we require a method to rotate the planar surface back to its original orientation.
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=3.5cm]{had11.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=3cm]{had12.jpg}
}
\caption{Performing a transversal Hadamard gate leaves the original qubit surface (a) in a rotated orientation (b).}
\label{had1}
\end{figure}
We can perform this rotation using the method shown in figure \ref{had2}. The original surface in figure \ref{had1}(b) is expanded to create the surface in figure \ref{had2}(a). This is still a planar surface with two sets of boundaries, with the smallest logical operator strings between boundaries of length $d$ (the original code distance). Applying $d$ rounds of error correction after this merging maintains fault tolerance. The large surface is then contracted by measuring out qubits in the $Z$ basis, figure \ref{had2}(b). After a further $d$ rounds of error correction this leaves the remaining surface correctly oriented for further interactions with other planar code surfaces, but shifted by half a lattice spacing in both horizontal and vertical directions. This can be corrected by performing swap operations to move the lattice into the correct position.
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=5cm]{had21.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=5cm]{had22.jpg}
}
\caption{Rotating the orientation of a planar qubit: a) expanding the original surface (pink); b) contracting to form the rotated surface (pink).}
\label{had2}
\end{figure}
\section{Relationship to defect qubits}\label{relate}
We saw in \S\ref{scs} that there is a close relationship between logical qubits defined with respect to the boundaries of the lattice, and those defined by the ``extra boundaries" of double defects. We can in fact convert between the two types, extruding a defect-based qubit from the edge of a surface into a planar qubit on a separate surface. As well as demonstrating the connection between the two surface code types, such a procedure could be useful if, for example, single qubits need to be extracted from a larger computation and then distributed. Reducing the number of physical qubits to be communicated in this case would be very useful.
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=7cm]{deftoplan21.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=7cm]{deftoplan22.jpg}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=7.5cm]{deftoplan23.jpg}
}
\caption{Extracting a planar code qubit from a double defect qubit located at the edge of a larger lattice: a) pink qubits are measured out in the $Z$ basis to isolate the defect qubit from the rest of the lattice; b) the defects are enlarged to fill part of the space; c) unstabilized qubits are removed from the lattice, creating the planar qubit. Note that this procedure also works in reverse.}
\label{extrude}
\end{figure}
The full procedure, complete with correction operations at the boundary, is detailed in \cite{austinH}. Let us start with a logical qubit defined by a double defect, near the edge of a large lattice, figure \ref{extrude}(a). We can produce a planar qubit from this defect qubit as in figure \ref{extrude}, where the pink qubits are removed from the lattice by measurements. In this case, with a smooth double defect, the measurements are in the $Z$ basis. If the defects were rough, the measurements would need to be in the $X$ basis. This leaves us with an isolated planar qubit, which is no longer attached to the main surface (and can indeed be detached for communication), and encodes the same logical qubit (at the same code distance) as the original defect-based logical qubit.
\section{Resource use of the planar code}\label{resource}
Figure \ref{distfour} shows clearly the significant difference in resources used for planar and defect-based logical codes of the same distance. We can now quantify this difference for individual logical qubits, and also for entangling gates between them.
For both defect and planar qubits, full fault tolerance requires $d$ rounds of error correction after each operation on a distance $d$ code; operations on a single qubit therefore need the same number of time steps in both cases. However, if we look at the construction of the \leavevmode\hbox{\footnotesize{CNOT }} operation given in \S\ref{cnot}, it appears that a planar \leavevmode\hbox{\footnotesize{CNOT }} requires many more rounds of error correction than in a defect-based approach. In fact, if we look closely at the exact operations in the procedure as given, we find that this is not the case. Na\"ively, we would count operations for the steps as: first merge, 3 rounds; split, 3 rounds; second merge, 3 rounds. However, if we look again at the exact procedure as given is \S\ref{cnot}, we find that we do not in fact need the full 9 rounds of error correction. If we prepare the control qubit as a $d \times 2d$ surface, then the first merge operation is not required. Furthermore, all the required operations for the split and the merge commute; they can be performed at the same time, and then three rounds of error correction implemented afterwards. By thus combining the split and second merge, and eliminating the first merge, we can perform the logical \leavevmode\hbox{\footnotesize{CNOT }} with only $d$ rounds of error correction; the same as in the defect-based approach.
As we can make the temporal resources for each implementation equivalent in this way, the figure of merit to consider when comparing defect-based and planar implementations is going to be the number of physical qubits required -- that is, the lattice cross-section or surface area. For a distance $d$ code, a single double-defect qubit is usually implemented using a total lattice area of $\approx 6d^2$ qubits \cite{simon_architecture}. For a planar qubit, the lattice area is $\approx 2d^2$ qubits -- a significant reduction in physical qubits needed.
Let us now consider the requirements for the \leavevmode\hbox{\footnotesize{CNOT }} operation. The double defect implementation uses a large lattice cross-section to allow for the multiple braids that are required. The cross section used is $ \approx 37d^2$ qubits. For the planar code, by contrast, we use three $2d^2$ surfaces, and two lines of $2d$ intermediate data qubits, the equivalent of a single surface of cross section $2d(3d + 2)$. To leading order in $d$, then, this is a total reduction in the number of physical qubits used of around 6 times. We do not claim that the current implementations of the defect-based code are optimal. The straightforward calculation is not, however, the whole story, as we need also in general to consider the layout of qubits that will allow for scalable computation. The numbers given here for the double-defect case are for a layout that is designed to scale arbitrarily. An equivalent layout for planar qubits is shown in figure \ref{scalableplanar}. The ``blank" areas of the lattice are set aside for \leavevmode\hbox{\footnotesize{CNOT }} operations, in order to allow any qubit to perform a \leavevmode\hbox{\footnotesize{CNOT }} with any other qubit in the computer. Only a quarter of the available surfaces can then be used to hold logical data qubits -- the rest must remain available for logical operations.
\begin{figure}[t]
\centering
\includegraphics[width=4cm]{scalableplanar}
\caption{Scalable arrangement of planar code surfaces. Shaded surfaces contain logical data qubits, and blank surfaces are available to perform \leavevmode\hbox{\footnotesize{CNOT }} operations between data surfaces.}\label{scalableplanar}
\end{figure}
We can therefore see that, for a large-scale quantum computer, there is very little difference in the resource requirements for planar or defect implementations. The area in which there is a significant difference is where the number of operations and qubits considered is small. For medium-scale processes, with tens of physical qubits, the above comparisons hold. However, if we look at small scale surface code implementations, the appropriate comparison will not be with the double-defect code, but rather the use of a single defect per logical qubit. For the single-defect implementation, a single logical qubit in the best-known implementations uses a surface area of $ \approx 10d^2$ qubits. When considering a single \leavevmode\hbox{\footnotesize{CNOT }} operation, this can be performed by looping a rough defect (created by not enforcing a vertex stabilizer) all the way around a smooth defect (created by not enforcing a face stabilizer), and uses only the same surface area to leading order in $d$. Exact calculations of the size of the surface depend on the exact code distance and the order of the boundaries used, but the single-defect implementation usually uses between 1.5 and 2 times the qubits of the planar implementation. As we go to smaller scales, and start to make contact with experimentally feasible implementations of the surface code, this difference can become significant.
\section{Small scale experiments on the planar code}
One of the motivations for introducing lattice surgery on the planar surface code was the reduction in qubit requirements for small scale implementations and medium-term achievable experiments. We have seen that the planar implementation indeed uses fewer resources than either the single or double defect model. We now look at exactly how small the planar code can go, and still provide useful experimental results. In order to do this, we first outline a useful modification to the surface code where the qubit lattice is rotated. We then give two proposed medium-term achievable experiments. In the first, lattice surgery is used to create entangled qubits, both Bell pairs and GHZ states. In the second, we give the precise requirements and procedure for the smallest non-transversal planar \leavevmode\hbox{\footnotesize{CNOT }}. In both cases we find that the addition of lattice surgery to the surface code toolkit significantly reduces the resources required to perform these error correction experiments.
\subsection{The rotated lattice}
The standard method for creating planar encoded qubits uses the square lattice with regular boundaries, as in figure \ref{distfour}(a). However, it is possible to reduce the number of physical qubits required for a single planar surface of a given distance by considering a ``rotated" form of the lattice used. This removes physical qubits from the edges of the lattice, creating irregular boundaries. The shortest string of operators creating a logical operator is frequently then not a straight line; however, as we shall see, it never goes below the code distance of the original surface.
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=4.5cm]{5x5}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=3cm]{5x5-rot1}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=5.5cm]{5x5-rot2}
}
\caption{Rotating a distance 5 lattice to produce another distance 5 encoded qubit. a) The original surface. The red square shows the area of the rotated surface. b) The rotated surface; (c) the rotated plaquettes (vertex plaquettes are marked brown, face plaquettes as yellow), new example logical operators, and boundaries. $X$($Z$) logical operators are shown as red(blue) lines.}
\label{5x5lattices}
\end{figure}
Let us consider the distance 5 code surface shown in figure \ref{5x5lattices}(a). We now create a ``rotated" lattice form by removing all the qubits outside the red box, and rotating (for clarity) $45^\circ$ clockwise. This now has the logical operators and boundaries as shown in figure \ref{5x5lattices}(b). If we now colour in the stabilizers in this new rotated form, we have the lattice in \ref{5x5lattices}(c), with 25 physical qubits and 24 independent stabilizers. The boundaries are now no longer ``rough" or ``smooth". Instead we use \emph{X-boundaries} and \emph{Z-boundaries}: $X$-boundaries have $X$ syndrome measurements along the boundary (shown in the figures as brown), and $Z$-boundaries have Z-syndrome measurements along the edge (shown as yellow) \cite{planar-bk}.
There are now more possible paths across the lattice area for a logical operator to take; the smallest paths do, however, stay the same length as before the rotation, so the rotated and unrotated lattices have the same error correction strength. Figure \ref{5x5lattices}(c) shows several example logical operators; note that, in the rotated form, logical operator chain paths can go diagonally through plaquettes of the opposite type (i.e.. $X$-chains can go vertically through $Z$-plaquettes, and \emph{vice versa}). Not all such paths are possible logical operators: care must be taken to ensure that each face(vertex) plaquette is touched an even number of times for an $X$($Z$) operators. This maintains the commutation of the operator with the lattice stabilizers, and the logical operator operations are undetectable by syndrome measurements. The code distance of the logical qubit remains unchanged, and we have reduced the number of lattice data qubits from $d^2 + (d-1)^2$ to $d^2$ for a distance $d$ code.
\begin{figure}[t]
\centering
\subfigure[]
{\includegraphics[width=4.5cm]{3x3}
}\hspace{1cm}
\subfigure[]
{\includegraphics[width=4.5cm]{3x3-rotated}
}\hspace{1cm}
\caption{Two lattices encoding a single qubit with distance 3: a) standard planar lattice; b) the rotated lattice. Light(dark) plaquettes show $Z(X)$ syndrome measurements.}
\label{twolattices}
\end{figure}
The smallest code that detects and corrects one error is the distance 3 code, figure \ref{twolattices}(a). The rotated lattice is shown in figure \ref{twolattices}(b). Note that measuring all the boundary syndromes is unnecessary as the stabilizers are not independent; only the syndromes marked are measured. We can explicitly demonstrate the action of creating the rotated lattice by considering all the stabilizers of the surface, choosing the logical state to be the +1 eigenstate of the $X_L$ logical operator. The left-hand column of figure \ref{stabsrotorig} shows the surface stabilizers of the standard lattice in this case, and the right-hand column gives the corresponding stabilizers in the rotated case. The `rotated' encoding allows us to produce a distance 3 planar qubit with 9 data qubits and 8 syndromes that require measurement. This can be done with either 8 syndrome qubits, or else the four central syndrome qubits can be used twice (while still requiring only neighbouring qubits to interact). The rotated encoding therefore reduces the number physical qubits for the smallest code distance from 25 to 13.
\begin{figure}[t]
\centering
\begin{tabular}{c || c}
Standard lattice & Rotated lattice\\
stabilizers & stabilizers\\ \hline
$X_2X_7X_{12} \ (=X_L)$ & $X_2X_7X_{12} \ (=X_L)$ \\
$X_1X_2X_4$ & {}\\
$X_2X_3X_4$ & $X_2 X_4$\\
$X_4X_6X_7X_9$ & $X_4X_6X_7X_9$\\
$X_5X_7X_8X_10$ & $X_5X_7X_8X_{10}$ \\
$X_9X_{11}X_{12}$ & {}\\
$X_{10}X_{12}X_{13}$ & $X_{10}X_{12}$\\
$Z_1Z_4Z_6$ & {}\\
$Z_2Z_4Z_5Z_7$ & $Z_2Z_4Z_5Z_7$\\
$Z_3Z_5Z_8$ & $Z_5 Z_8$\\
$Z_6Z_9Z_{11}$ & $Z_6Z_9$\\
$Z_7Z_9Z_{10}Z_{12}$ & $Z_7Z_9Z_{10}Z_{12}$\\
$Z_8Z_{10}Z_{13}$ & {}\\
\end{tabular}
\caption{Stabilizers for the distance 3 planar qubit of figure \ref{twolattices}, encoding the state $\ket{+}$, for the standard and rotated lattices. Note that the vertical chain $X_4X_7X_{10}$ is also a valid logical $X_L$ operator. } \label{stabsrotorig}
\end{figure}
\subsection{Experiment 1: creating entangled states}
We can use lattice surgery, specially the splitting operation, to generate entangled logical Bell pairs and GHZ states. This has an advantage over other methods of preparing entangled logical states that more complicated 2-qubit logical operators are not required. The states created can be n-dimensional GHZ states in either the computational or Hadamard basis.
\begin{figure}[t]
\centering
\includegraphics[width=12cm]{GHZ-rot.jpg}
\caption{Creating a logical GHZ state on three surfaces in the computational basis through lattice splitting. The orange lattice denotes the rotated encoding.}\label{ghz}
\end{figure}
The procedure is shown in figure \ref{ghz} for distance 4 logical qubits. The initial continuous surface is prepared in the logical $\ket{+}$ state, the +1 eigenstate of the $X_L$ logical operator. The surface is then split by measuring one row of the pink qubits -- this is a smooth split, so the resultant state is
\begin{equation} \ket{+} = \frac{1}{\sqrt{2}}( \ket{0} + \ket{1} ) \longrightarrow \frac{1}{\sqrt{2}}( \ket{00} + \ket{11} )\end{equation}
\noindent which is our Bell pair. If we then measure out the second pink qubit line, we perform another smooth split taking us to a 3-qubit GHZ state:
\begin{equation} \frac{1}{\sqrt{2}}( \ket{00} + \ket{11} ) \longrightarrow \frac{1}{\sqrt{2}}( \ket{000} + \ket{111} )\end{equation}
It is trivial to see that subsequent splittings will produce higher-dimensional GHZ states. The only constraint is that the initial surface is large enough to produce individual logical qubits at the end of the splitting process that retain the desired code distance. We can see that GHZ states in the Hadamard basis can be produced in exactly the same way, with rough splitting. The shape of the original surface will require modification to allow the produced qubits to support the required code distance.
\subsection{Experiment 2: the smallest planar \leavevmode\hbox{\footnotesize{CNOT }}}
\begin{figure}[t]
\centering
\includegraphics[width=8cm]{sd-cnot-stabilizers.jpg}
\caption{A standard \leavevmode\hbox{\footnotesize{CNOT }} operation between two logical defects of distance 3 defined by single defects on a lattice. The pink defect is moved around the green one to complete the gate.}\label{cnot-sd}
\end{figure}
The distance 3 rotated planar lattice can also be used to define the smallest number of physical qubits required for a single logical \leavevmode\hbox{\footnotesize{CNOT }} operation, using the lattice surgery methods of \S\ref{cnot}. With the usual surface code methods, the restriction to nearest-neighbouring-only interactions means that the smallest logical \leavevmode\hbox{\footnotesize{CNOT }} requires relatively large numbers of physical qubits to implement. Using a double defect, the best-optimised \leavevmode\hbox{\footnotesize{CNOT }} currently known uses 143 qubits. Using single defects, we have the lattice in figure \ref{cnot-sd}, which also uses 143 physical qubits. By implementing this on a rotated lattice, this can be reduced further to 104 qubits.
\begin{figure}[t]
\centering
\includegraphics[width=8cm]{cnot-rot-STABILIZERS.jpg}
\caption{Lattice qubits for a lattice surgery logical \leavevmode\hbox{\footnotesize{CNOT }} between two distance 3 logical qubits, using the rotated encoding (lattice shown is for the standard encoding to allow comparison with figure \ref{cnot-sd}).}\label{cnot-small}
\end{figure}
The surgery method reduced this number still further. The arrangement of data qubits is shown in figure \ref{cnot-small}. Split and merge operations are performed between the control and target using the method given in \S\ref{cnot}. By using the rotated encoding, and eliminating syndrome qubits outside the data qubit lattice, we require 33 data qubits and 20 syndrome qubits, for a total of 53 physical qubits. We can therefore halve the number of physical qubits required for the smallest \leavevmode\hbox{\footnotesize{CNOT }} gate implementation by using lattice surgery. While not implementable using current experimental techniques, this number of qubits is by no means out of the question for implementation over the next few years.
\section{Conclusions}
In this paper we have introduced a method for interacting multiple logical qubits encoded with the planar code, without the need to break the 2DNN structure of the error correction. This procedure, which we term lattice surgery, allows us to perform coupling operations between encoded qubits without utilising standard transversal protocols. This method maintains both fault-tolerance and universality while ultimately reducing qubit resources.
While qubit resource savings are comparatively modest, any saving is advantageous for short to medium term experiments and prototype systems. The ability of planar codes to be coupled without a full transversal protocol will also be beneficial in the context of distributed computation and/or communication. As lattice surgery only requires interaction along boundaries, the number of distributed Bell state necessary to perform logical operations between distributed planar qubits is reduced, relaxing the requirements of a repeater network responsible for the connections.
In addition to the universal set of gates required for large-scale implementation, we have also illustrated some more basic operations: the generation of encoded Bell and GHZ states, and the smallest possible logical operation that can be performed on topologically protected qubits designed to correct for a single arbitrary error on each logical qubit space. These operations will undoubtably be the first to be experimentally demonstrated once qubit technology reaches the level of 50-100 qubits.
\section{Acknowledgements}
The authors would like to thank Ashley Stephens and David DiVincenzo for comments on the manuscript. CH, RV and SD acknowledge that this research is supported by the JSPS through its FIRST Program. S.D acknowledges support from MEXT. AGF acknowledges support from the Australian Research Council Centre
of Excellence for Quantum Computation and Communication Technology
(Project number CE110001027), and the US National Security Agency
(NSA) and the Army Research Office (ARO) under contract number
W911NF-08-1-0527.
\section*{References}
\bibliographystyle{unsrt}
|
\section{Introduction and motivation}\label{sec:int}
We consider a family of equivalence relations on permutations in
$S_{n}$ in which two permutations are considered to be equivalent if
one can be converted into the other by replacing a short subsequence of
(not necessarily adjacent) elements by the same elements permuted
in a specific fashion, or (extending by transitivity) by a sequence
of such moves. These generalise the relations discovered by Knuth
in his study of the Robinson-Schensted correspondence, though the
original motivations for this project were unrelated. We begin the
systematic study of such equivalence relations, connecting them with
integer sequences both familiar and (apparently) new.
Consider the following three examples of turning one 7-permutation into another
in which selected 3-subsequences (marked in \textbf{bold}) are re-ordered:
\begin{eqnarray}\label{eqn:eg1}
12\mathbf{34}56\mathbf{7} &\rightarrow & 12\mathbf{74}56\mathbf{3}\\
\mathbf{127}4563 &\rightarrow & \mathbf{721}4563\\
721\mathbf{456}3 &\rightarrow &721\mathbf{654}3
\end{eqnarray}
In each of these examples, a subsequence of \textit{pattern} 123 (i.e.,
a triple of not necessarily adjacent entries whose elements are in the
same relative order as 123) is replaced by the same set of elements arranged
in the pattern 321. Allowing replacements of a designated kind to be
performed ad libitum, in reverse as well as forward, induces an equivalence
relation on the symmetric group $S_{7}$. Accordingly we can say that the
permutations 1234567, 1274563, 7214563, and 7216543 are all equivalent
under the replacement $123 \leftrightarrow 321$.
Interesting enumerative questions arise when the elements being replaced
are allowed to be in general position (Section~\ref{sec:gen}), when the
replacements are constrained to involve only adjacent elements
(Section~\ref{sec:adj}), and when replacements are constrained to affect
only subsequences of consecutive elements representing a run of
consecutive values (Section~\ref{sec:dbl}). Each of these respective
\emph{types} of replacements is illustrated in one of the three examples
above. It will be convenient to group subsequences that are allowed to
replace one another into sets, e.g., describing the three permutations
above as being ``$\{123,321 \}$-equivalent''. We may also wish to allow
more than one type of (bi-directional) replacement, such as both $123
\leftrightarrow 321$ and $123 \leftrightarrow 132$. If the intersection
of these sets is nonempty, the new relation can be described simply by
the union of the two sets: $\{123,132,321\} = \{123,321\} \cup
\{123,132\}$. To formalise this more generally we consider collections
of disjoint replacement sets that form a set partition of $S_{3}$; any
two patterns within the same set may replace one another within the
larger permutation to give an equivalent permutation.
Let $\pi \in S_{n}$, and let $P=\{B_{1},B_{2},\dots,
B_{t}\}$ be a (set) partition of $S_{k}$, where $k\leq n$. Each block
$B_{l}$ of $P$ represents a list of $k$-length patterns which can
replace one another. We call two permutations $P^{\fourdots}$-equivalent if one
can be obtained from the other by a sequence of replacements, each
replacing a subsequence of pattern $\sigma_{i}$
with the same elements in the pattern $\sigma_{j}$,
where $\sigma_{i}$ and $\sigma_{j}$ lie in the same block
$B_{l}$ of $P$. Let $\mathrm{Eq}^{\fourdots}(\pi,P)$ denote the set of permutations
equivalent to $\pi$ under $P^{\fourdots}$-equivalence; e.g.,
1234567, 7214563, and $7216543\in
\mathrm{Eq}^\fourdots\left(1274563,\big\{\{123,321\}\big\}\right)$.
Similarly we denote by $P^{\twolines}$ the equivalence
relation, and by $\mathrm{Eq}^{\twolines}(\pi,P)$ the equivalence
class of $\pi$, under replacement within $P$ only of adjacent elements; e.g.,
7214563 and $7216543 \in \mathrm{Eq}^\twolines\left(1274563,\big\{\{123,321\}\big\}\right)$.
We use $P^{\square}$ and $\mathrm{Eq}^\square(\pi,P)$ when both positions and values are
constrained, e.g.,
$7214563 \in \mathrm{Eq}^\square\left(7216543,\big\{\{123,321\}\big\}\right)$.
To refer to such classes generically we use the notation
$\mathrm{Eq}^*(\pi,P)$. The automorphism $\pi\mapsto \pi^{-1} $ replaces
adjacency of positions with adjacency of values; hence, for the enumerative
questions we treat, there is no need to separately consider a fourth
case where we only constrain values to be adjacent.
The set of distinct equivalence classes into which $S_n$ splits under
an equivalence $P^{*}$ is denoted by $\mathrm{Classes}^*(n,P)$.
The present paper begins the study of these equivalence relations by
considering three types of questions:
(A) Compute the number of equivalence classes,
$\#\mathrm{Classes}^{*}(n,P)$, into which $S_{n}$ is partitioned.
(B) Compute the size, $\#\mathrm{Eq}^*(\iota_n,P)$, of the equivalence
class containing the identity $\iota _{n}=123\cdots n$.
(C) Characterise the set $\mathrm{Eq}^*(\iota_n,P)$ of
permutations equivalent to the identity.
Although the framework above allows for much greater generality, in this
paper we will mainly restrict our attention to replacements by patterns of
length $k=3$, and usually to replacement patterns built up from pairs in
which one permutation is the identity, and the other is a transposition
(i.e., fixes one of the elements). Omitting some cases by symmetry,
we have the following
possible partitions of $S_{3}$, where (as usual) we omit singleton blocks:
\begin{eqnarray*}
P_1 = \big\{ \{123, 132\} \big\}, \\
P_2 = \big\{ \{123, 213\} \big\}, \\
P_4 = \big\{ \{123, 321\} \big\}.
\end{eqnarray*}
We will also consider applying two of these replacement operations
simultaneously, and we will number the appropriate partitions as
\begin{eqnarray*}
P_3 = \big\{ \{123, 132, 213\} \big\}, \\
P_5 = \big\{ \{123, 132, 321\} \big\}, \\
P_6 = \big\{ \{123, 213, 321\} \big\},
\end{eqnarray*}
following the convention $P_{i+j} := P_i \vee P_j$, the \emph{join} of
these two partitions~\cite[ch.3]{StanleyEC1}. Indeed we
can allow all three replacements: $P_7 = \big\{ \{123, 132, 213, 321\} \big\}$.
(In fact, the cases $P_1$ and $P_2$ are equivalent by symmetry, as
are $P_5$ and $P_6$. We list $P_1$ and $P_2$ separately only
in order to consider their join.)
Our motivation for focussing attention on pairs of this form is that
we can then think of an operation, not in terms of replacing one
pattern by another, but simply in terms of \emph{swapping} two
elements within the pattern, with the third serving as a catalyst
enabling the swap. In a followup~\cite{PRW11} to the current paper, the
authors treat the remaining (non-swapping) cases for all partitions of
$S_{3}$ consisting of exactly one non-singleton block which contains the
identity $123$.
By far the best-known example of constrained swapping in permutations is
the \emph{Knuth Relations}~\cite{Knuth}, which allow the swap of
adjacent entries provided an intermediate value lies immediately to the
right or left. In the notation of this paper, they correspond to
$P^{\twolines}_{K}=\big\{ \{ 213,231\}, \{132,312\}\big\}$. Permutations
equivalent under this relation map to the same first coordinate
($P$-tableau) under the Robinson-Schensted correspondence.
Mark Haiman introduced the notion of \emph{dual equivalence} of
permutations: $\pi$ and $\tau$ are dual equivalent if one can be
obtained from the other by swaps of adjacent \emph{values} from the above
$P_{K}$, i.e., if their inverses are Knuth-equivalent, or
(equivalently) if they map to
the same second coordinate ($Q$-tableau) under the Robinson-Schensted
correspondence~\cite{HaimanDEq}. For the enumerative problems in this
paper, we get the same answers for Knuth and dual equivalence.
In her dissertation~\cite{AssafDEG} Sami Assaf constructed graphs (with
some extra structure) whose vertices are tableaux of a fixed shape
(which may be viewed as permutations via their ``reading words''), and
whose edges represent (elementary) dual equivalences between
vertices. For this particular relation (equivalently for the Knuth
relations), she was able to characterise the local structure of these
graphs, which she later used to give a combinatorial formula for the Schur
expansion of LLT polynomials and MacDonald Polynomials. She also used
these, along with crystal graphs, to give a combinatorial realization of
Schur-Weyl duality~\cite{AssafSWD}.
Sergey Fomin has a very clear elementary exposition of how Knuth and dual
equivalence are related to the Robinson-Schensted correspondence,
Sch\"utzenberger's \textit{jeu de taquin}, and the Littlewood-Richardson
rule in~\cite[Ch.~7, App.~1]{StanleyEC2}. For the problems considered
above, the answers for $P_{K}^{\twolines}$ are well known to be: (A) the
number of involutions in $S_{n}$; (B) 1; and (C) $\{\iota_{n}\}$. In fact one
can compute $\#\mathrm{Eq}^{\twolines}(\pi ,P_{K})$ for any permutation $\pi$ by
using the Frame-Robinson-Thrall hook-length formula to compute the
number of standard Young tableaux of the \emph{shape} output by the
Robinson-Schensted correspondence applied to $\pi$.
Any of the relations we consider can be naturally generalized to operate
on $\mathcal{W}([n])$, the set of \textbf{words} (with repeated entries
allowed) on the alphabet $[n]$: for example, the relation $123
\leftrightarrow 321$ would imply also moves of the form
$112\leftrightarrow 211$ and $122\leftrightarrow 221$, treating letters
with the same label within a word as increasing from left to right. In
the case of the Knuth relations, the equivalence classes are simply the
\emph{elements} of the well-known \textit{plactic monoid} of Lascoux and
Sch\"utzenberger: $\mathcal{W}([n])/P_{K}$ \cite{LSPlactic, LLTPlactic}.
In \cite{Chinese}
the authors study the analogous
\textit{Chinese monoid}, which is $\mathcal{W}([n])/P_{3}$ (up to
the involution that reverses all words), for which they develop an
analogue of the Robinson-Schensted correspondence and count some of the
equivalence classes.
Given that the Knuth relations act on adjacent elements, and lead to
some deep combinatorial results, it is perhaps not surprising that the
most interesting problems and proofs in this paper are to be found in
Section~\ref{sec:adj}. A summary of our numbers and results is given
in Figure~\ref{fig:sum}.
An extended abstract of this paper appeared in the proceedings of
FPSAC10~\cite{LPRW}. The third author is grateful to Sami Assaf, Karen
Edwards and Stephen Pon for helpful conversations.
\begin{figure}[h]
\caption{Summary of Results}
\label{fig:sum}
\medskip
\begin{minipage}[t]{\textwidth}
These tables give numerical values and names (when available) of the
sequences associated with enumerative questions (A) and (B). All
sequences begin with the value for $n=3$. Results proven in this paper
have a gray background; for other cases we lack even conjectural
formulae. Six-digit codes preceded by ``A'' cite specific sequences in
Sloane~\cite{OEIS}.
\end{minipage}
\medskip
\begin{tiny}
\begin{large}Number of classes\end{large}
\bigskip
\begin{tabular}{|c|l|l|l|l|}
\hline
\multicolumn{2}{|l|}{\multirow{2}{*}{\normalsize Transpositions}} & \small \S~2 & \small \S~3& \small\S~4 indices and\\
\multicolumn{2}{|l|}{} &{\small general} & {\small only indices adjacent} & {\small values adjacent} \\
\hline
(1) & $123 \leftrightarrow 132$ & \cellcolor{lightgray}[5, 14, 42, 132, 429] &\multirow{2}{*}{[5, 16, 62, 284, 1507, 9104]} & \multirow{2}{*}{[5, 20, 102, 626, 4458, 36144]} \\
\cline{1-2}
(2) & $123 \leftrightarrow 213$ & \cellcolor{lightgray}Catalan & & \\
\hline
\multirow{2}{*}{(4)} & \multirow{2}{*}{$123 \leftrightarrow 321$} & \cellcolor{lightgray}[5, 10, 3, 1, 1, 1] & \multirow{2}{*}{[5, 16, 60, 260, 1260, 67442] } & \multirow{2}{*}{[5, 20, 102, 626, 4458, 36144]} \\
& & \cellcolor{lightgray}trivial & & \\
\hline
\multirow{2}{*}{(3)} & \multirow{2}{*}{$123 \leftrightarrow 132 \leftrightarrow 213$} & \cellcolor{lightgray}[4, 8, 16, 32, 64, 128] & \cellcolor{lightgray}[4, 10, 26, 76, 232, 764] & \multirow{2}{*}{[4, 17, 89, 556, 4011, 32843]} \\
& & \cellcolor{lightgray}powers of 2 &\cellcolor{lightgray}involutions &\\
\hline
(5) & $123 \leftrightarrow 132 \leftrightarrow 321$ & \cellcolor{lightgray}[4, 2, 1, 1, 1, 1] & \multirow{2}{*}{[4, 8, 14, 27, 68, 159, 496]} & \multirow{2}{*}{[4, 16, 84, 536, 3912, 32256]} \\
\cline{1-2}
(6) & $123 \leftrightarrow 213 \leftrightarrow 321$ & \cellcolor{lightgray}trivial & & \\
\hline
\multirow{2}{*}{(7)} & $123 \leftrightarrow 132$ & \cellcolor{lightgray}[3, 2, 1, 1, 1, 1] & \multirow{2}{*}{[3, 4, 5, 8, 11, 20, 29, 57]} & \multirow{2}{*}{[3, 13, 71, 470, 3497]} \\
& $\leftrightarrow 213 \leftrightarrow 321$ & \cellcolor{lightgray}trivial & &\\
\hline
\end{tabular}
\vspace{3mm}
\begin{large}Size of class containing identity\end{large}
\bigskip
\begin{tabular}{|c|l|l|l|l|}
\hline
\multicolumn{2}{|l|}{\multirow{2}{*}{\normalsize Transpositions}} & \small \S~2 & \small \S~3& \small\S~4 indices and\\
\multicolumn{2}{|l|}{} &{\small general} & {\small only indices adjacent} & {\small values adjacent} \\
\hline
(1) & $123 \leftrightarrow 132$ &\cellcolor{lightgray}[2, 6, 24, 120, 720] &\cellcolor{lightgray}[2, 4, 12, 36, 144, 576, 2880] & \cellcolor{lightgray}[2, 3, 5, 8, 13, 21, 34, 55] \\
\cline{1-2}
(2) & $123 \leftrightarrow 213$ & \cellcolor{lightgray}$(n-1)!$ &\cellcolor{lightgray}product of two factorials & \cellcolor{lightgray}Fibonacci numbers \\
\hline
\multirow{2}{*}{(4)} & \multirow{2}{*}{$123 \leftrightarrow 321$} &\cellcolor{lightgray}[2, 4, 24, 720] &\cellcolor{lightgray}[2, 3, 6, 10, 20, 35, 70, 126] & \cellcolor{lightgray}[2, 3, 4, 6, 9, 13, 19, 28] \\
& &\cellcolor{lightgray}trivial & \cellcolor{lightgray}central binomial coefficients & \cellcolor{lightgray}A000930 \\
\hline
\multirow{2}{*}{(3)} & \multirow{2}{*}{$123 \leftrightarrow 132 \leftrightarrow 213$} &\cellcolor{lightgray}[3, 13, 71, 461] & [3, 7, 35, 135, 945, 5193] &\cellcolor{lightgray}[3, 4, 8, 12, 21, 33, 55, 88] \\
& &\cellcolor{lightgray}connected A003319 & Chinese Monoid~\cite{Chinese} &\cellcolor{lightgray}A052952 \\
\hline
(5) & $123 \leftrightarrow 132 \leftrightarrow 321$ &\cellcolor{lightgray}[3, 23, 120, 720]
&\cellcolor{lightgray}[3, 9, 54, 285, 2160, 15825] &\cellcolor{lightgray}[3, 5, 9, 17, 31, 57, 105, 193] \\
\cline{1-2}
(6) & $123 \leftrightarrow 213 \leftrightarrow 321$ &\cellcolor{lightgray}trivial &\cellcolor{lightgray}separate formulae for odd/even &\cellcolor{lightgray}tribonacci numbers A000213 \\
\hline
\multirow{2}{*}{(7)} & $123 \leftrightarrow 132$ & \cellcolor{lightgray}[3, 23, 120, 720] &[4, 21, 116, 713, 5030] &\cellcolor{lightgray}[4, 6, 13, 23, 44, 80, 149, 273] \\
& $\leftrightarrow 213 \leftrightarrow 321$ & \cellcolor{lightgray}trivial &
&\cellcolor{lightgray}tribonacci A000073 $-[n\ \mathrm{ even}]$
\\
\hline
\end{tabular}
\end{tiny}
\end{figure}
If $\tau \in \mathrm{Eq}^*(\pi,P)$ we will say that $\tau$ is \emph{reachable}
from $\pi$ (under $P$). If $\mathrm{Eq}^*(\iota_n,P) = S_n$, then every
permutation in $S_n$ is reachable from every other, and we will say that
$S_n$ is \emph{connected} by $P$. If $\mathrm{Eq}^*(\pi,P) =\{\pi\}$
we will say that $\pi$ is \emph{isolated} (under $P$).
It is obvious that if $P_i$ refines $P_j$ as partitions of $S_{k}$
(i.e., $P_{i}\leq P_{j}$ in the lattice of partitions of $S_{k}$),
then the partition of $S_n$ induced by $P_i$ refines the one induced by
$P_j$, because a permutation reachable from $\pi$ under $P_i$ is also
reachable under $P_j$. This enables the following simple observations:
\begin{propo}\label{propo:Refine}
If $P_i$ refines $P_j$ (as partitions of $S_{k}$), then for all $\pi \in
S_{n}$ with $n\geq k$
\begin{eqnarray*}
\mathrm{Eq}^*(\pi,P_i) \subseteq \mathrm{Eq}^*(\pi,P_j) \\
\#\mathrm{Eq}^*(\pi,P_i) \leq \#\mathrm{Eq}^*(\pi,P_j) \\
\#\Classes^*(n,P_i) \geq \#\Classes(n,P_j)
\end{eqnarray*}
\end{propo}
\section{General pattern equivalence}\label{sec:gen}
In this section, we allow moves within an equivalence relation with no
adjacency restrictions. This case is closely related to the theory of
\textit{pattern avoidance\/} in permutations: replacing one
pattern with another repeatedly leads eventually to a permutation which
avoids the first pattern.
Some of the equivalence relations in this section are trivial, following
immediately from the following observation. The others lead to familar
combinatorial numbers and objects.
\begin{propo}\label{propo:OneClass}
Fix $k$ with $2\leq k\leq n-1$, and let $P$ be any partition of
$S_{k}$.
If $\#\Classes^{\fourdots}(n-1,P) = 1$, then $\#\Classes^{\fourdots}(n,P) = 1$.
\end{propo}
\begin{proof}
We will show that any $\pi \in S_n$ can be reached from the identity, $\iota_{n}$,
under the supposition that any two permutations in $S_{n-1}$ are
equivalent, in two stages. If $\pi(1) \neq n$, simply apply the
supposition to the elements/positions $1 \dots n-1$ in $\iota_{n}$ to
obtain any permutation beginning with $\pi(1)$; if $\pi(1)=n $, we use
instead the elements/positions $1,3,4,5,\dotsc n$ (omitting 2, which is
$\leq n-1$ by hypothesis) to move $\pi (1)=n$ to the front of a
permutation equivalent to $\iota_{n}$. Then in stage 2 we apply the
supposition to the elements now occupying \textit{positions\/} $2,
\dots, n$ to complete the construction of $\pi$.
\end{proof}
The following results follow.
\begin{propo}\label{propo:P457OneClass}
$\#\Classes^\fourdots\left(n,\big\{ \{123,321\} \big\}\right) = 1$ for $n \geq
6$. While for $n \geq 5$, we have $\#\Classes^\fourdots\left(n,\big\{ \{123,132,321\} \big\}\right) = 1$
and $\#\Classes^\fourdots\left(n,\big\{ \{123,132,213,321\} \big\}\right) = 1$
\end{propo}
\begin{proof}
It is easy to verify by hand, or by computer, that all permutations
in $S_5$ are reachable from $12345$ by moves in $P_5 = \big\{ \{123,132,321\} \big\}$.
(Indeed, all permutations in $S_4$ are reachable from $1234$ except for
$3412$, which is isolated.) As $S_5$ is connected, it follows
(by induction) from the preceding proposition that $S_n$ is connected
for all $n\geq 5$.
Proposition~\ref{propo:Refine} tells us that $S_n$ is connected under
$P_7 = \big\{ \{123,132,213,321\} \big\}$ whenever it is connected under
$P_5$ since $P_7 \geq P_5$. (In $S_4$, the permutation $3412$ remains
isolated.)
Finally, we can check by computer that under $P_4 = \big\{ \{123,321\} \big\}$
$S_6$ is connected; whence, $S_n$ is connected for $n \geq 6$.
\end{proof}
We remark that under $P_4$, $S_4$ splits into 10 equivalence classes,
and $S_5$ into three classes. The class containing $12345$ contains 24
elements. This suggests a possible bar bet. Hand your mark six cards
numbered $1$ through $6$ and invite him or her to lay them out in any
sequence. By applying moves of the form $123 \leftrightarrow 321$
(``Interchange two cards if and only if an intermediate (value) card
lies (in any position) between them.'') you will always be able to put
the cards in order. (It may take some practice, however, to become
proficient at doing this quickly.) Now ``go easy'' on your mark by
reducing the number of cards to 5. Even from a random sequence, the mark
has only one chance in five of being able to reach the identity.
Of course from Proposition~\ref{propo:P457OneClass} it immediately follows that:
\begin{cor}\label{cor:P457Id}
$\#\mathrm{Eq}^\fourdots\left(\iota_n,\big\{ \{123,132,321\} \big\}\right) =
n!$,
$\#\mathrm{Eq}^\fourdots\left(\iota_n,\big\{ \{123,132,213,321\} \big\}\right) = n!$ for $n \geq 5$; and
$\#\mathrm{Eq}^\fourdots\left(\iota_n,\big\{ \{123,321\} \big\}\right) = n!$ for $n \geq 6$.
\end{cor}
\begin{propo}\label{propo:P2Id}
$\#\mathrm{Eq}^\fourdots\left(\iota_n,\big\{ \{123,213\} \big\}\right) = (n-1)!$ for $n \geq 2$.
\end{propo}
\begin{proof}
Obviously the largest element $n$ cannot be moved away from the end
of the permutation. Equally obviously the $n$, remaining at the far right,
enables the other elements to be freely pairwise transposed, thereby
generating any permutation in $S_{n-1}$.
\end{proof}
\begin{propo}\label{propo:P2ClassCat}
For $n \geq 1$, $\#\Classes^\fourdots\left(n,\big\{ \{123,213\} \big\}\right) = c_n =
\frac{2n!}{n!(n+1)!}$, the $n$th Catalan number.
\end{propo}
\begin{proof}
Let $\pi \in S_{n}$. If $i < j < k$, and $\pi(i) < \pi(j) < \pi(k)$,
then $\pi(k)$ enables the swapping
of $\pi(i)$ and $\pi(j)$ to arrive at a permutation $\pi^{(1)}$ with a strictly larger
number of inversions. We can continue in this way to obtain a sequence
$\pi =\pi^{(0)}\rightarrow \pi^{(1)}\rightarrow \dots \rightarrow \pi^{(n)}$,
where $\pi^{(n)}$ has no such triples, i.e., $\pi^{(n)}$ is $123$-avoiding.
It remains to show that no matter which sequence of moves we make, the
final permutation $\pi^{(n)}$ is unique.
Call an element in a permutation $\sigma \in S_{n}$, a
\textit{right-to-left maximum} if it is greater than every element that
occurs to its right. (More formerly, $\sigma_{k}$ is a right-to-left
maximum if $\sigma_{k}>\sigma_{l}$ for all $k<l\leq n$.) The set
$\mathcal{M}(\pi )$ of these elements remains fixed under the relation
above, and forms a decreasing subsequence of $\pi^{(i)}$ for all $i$. Now
a permutation is $123$-avoiding if and only if it is a union of at most
two decreasing subsequences. So $\pi^{(n)}$ must be the unique
permutation formed by rearranging the elements of $\pi$ not in
$\mathcal{M}(\pi )$ in decreasing order. It is clear that the elements
of $\mathcal{M}(\pi )$ are positioned so as to enable all the necessary
transpositions.
Thus the ``largest'' (by number of inversions) elements in each equivalence
class are exactly the $123$-avoiding permutations, of which there are
$c_n$~\cite[ch.~14]{BonaWalk} or~\cite[Sec. 4.2]{BonaPerm}. (Similarly
one can show that the ``smallest'' elements are the $213$-avoiding
permutations.)
\end{proof}
\begin{eg}
\label{eg:P2ClassCat}
Working within $S_{9}$, we have the following sequence of equivalences,
where elements about to be transformed are indicated in \textbf{bold}.
The subsequence of right-to-left maxima is $976$.
\[
\mathbf{3}82941\mathbf{57}6 \rightarrow 58\mathbf{2}9\mathbf{4}13\mathbf{7}6\rightarrow
5849\mathbf{2}1\mathbf{37}6\rightarrow 58493\mathbf{127}6 \rightarrow
\mathbf{58}4\mathbf{9}32176 \rightarrow 854\mathbf{9}321\mathbf{76}
\]
The reader is encouraged to draw corresponding permutation matrices or
diagrams, which clarify visually how the right-to-left maxima facilitate
the transformation of the other elements into a decreasing subsequence.
\end{eg}
The next two propositions study an equivalence relation and class whose
enumeration is equivalent under symmetry (reversal or complementation)
to $\mathrm{Eq}^{\fourdots}(\iota_n,P_3)$.
The first leads to \textit{connected}
or \textit{indecomposable permutations}~\cite[A003319]{OEIS}, namely those not fixing
$\{1,2,\dotsc j\}$ for any $1\leq j <n$. If we define the
\textit{direct sum} of two permutations so that it corresponds to the
direct sum of the corresponding permutation matrices, then these are
simply the permutations which are indecomposable as direct sums in the
usual matrix sense. Some authors use the term
\textit{plus-indecomposable}~\cite{AAKSimple} to describe this class.
The second leads to the \textit{layered permutations}, namely those which are a
direct sum of decreasing permutations, introduced by
W.~Stromquist~\cite{StromLayered}, and studied carefully by A.~Price in
his thesis~\cite{PriceLayered}.
\begin{propo}\label{propo:P3revIndec}
Let $\rho_{n}$ denote the ``reverse word'' permutation $n,n-1,\dots 1$.
Then $\mathrm{Eq}^{\fourdots}(\rho_{n},\big\{ \{321,312,231\} \big\})$ is the set of indecomposable
permutations.
\end{propo}
\begin{proof}
When viewed as a $(0,1)$-matrix, any permutation decomposes as a direct sum of
irreducible blocks along the main diagonal; in particular, the identity
$\iota_n$ decomposes into $n$ singleton blocks, while $\rho_{n}$ is
indecomposable and is one large block. A permutation is connected if and
only if it decomposes as a single block.
First note that if any transformation of entries $(a_1,a_2,a_3) \rightarrow
(b_1,b_2,b_3)$ applied within a block causes it to split into more than
one block, then $b_1$ must be in the leftmost/lowest of the new blocks,
and $b_3$ in the rightmost/highest. Therefore $b_1$ must be less than
$b_3$, which is exactly what does not happen with any of our possible
transformations, because the first element is larger than the third in
each of 321, 312 and 231. Since all of our transformations are
reversible, this shows also that we cannot combine blocks. Thus,
\textit{the irreducible block structure of a permutation does not change
under these transformations.} In particular, if we start with an
indecomposable permutation such as $\rho_{n}$, successive applications
of the permitted operations will always produce indecomposable
permutations.
Next we have to show that all indecomposable permutations are in
fact reachable from $\rho_{n}$. Remembering that our replacement operations are
all reversible, we will instead show that we can always return to $\rho_{n}$
from an arbitrary indecomposable permutation.
Take $n\geq 3$, and let $\tau = \tau_1\tau_2\dotsb \tau_n$ be an arbitrary
indecomposable permutation other than $\rho_{n}$. We will show that $\tau $
always contains at least one of 312 or 231. It's easy to see that $\tau$
must have an \textit{ascent}, i.e., there exists $k$ such that
$\tau_{k}<\tau_{k+1}$.
Now if any element to the right of $\tau_{k+1}$ is less than $\tau_k$ we
have a 231, so assume there are none such. Similarly, assume there is no
element to the left of $\tau_k$ and greater than $\tau_{k+1}$ (avoiding
312).
But there must be some $y$ to the left of $\tau_k$ which is greater than
some $x$ to the right of $\tau_{k+1}$, or otherwise the permutation decomposes
between $\tau_k$ and $\tau_{k+1}$. These four elements $y,\tau_k,\tau_{k+1},x$ form a
3142, which contains both a 312 $(y,\tau_k,x)$ and a 231 $(y,\tau_{k+1},x)$.
Having now located a 312 or 231, we can then apply either $312 \rightarrow 321$
or $231 \rightarrow 321$, as appropriate. Each of these operations
simply switches
a pair of elements, and (as we have seen in the proof of
Proposition~\ref{propo:P2ClassCat}) strictly increases
the number of inversions, progressing us toward $\rho_{n}$.
This completes the proof that all indecomposable permutations are
reachable, and therefore the proof that the reachable permutations are
exactly the indecomposable permutations.
\end{proof}
\begin{propo}\label{propo:P3revNum}
$\#\Classes^\fourdots\left(n,\big\{ \{321,312,231\} \big\}\right) = 2^{n-1}$ for $n \geq 1$.
\end{propo}
\begin{proof}
As we saw in the proof of the previous proposition, the irreducible
block structure of a permutation does not change under the
transformations we are considering here. By the arguments already
given, we can work within any indecomposable block to restore it to an
anti-identity. Therefore each equivalence class consists of all the
permutations with a given block structure, and contains exactly one
permutation which is a direct sum of anti-identities.
These are exactly the layered permutations, and there are clearly
$2^{n-1}$ of them, with a factor of $2$ according to whether each consecutive
pair of elements is or is not in the same layer. (Equivalently, any such
permutation is determined by the composition of $n$ representing its
block sizes, of which there are $2^{n-1}$.)
\end{proof}
Finally we apply the reversal (or complementation) involution on $S_{n}$
to the above result to get our result for the partition $P_{3}$.
\begin{theorem}\label{thm:P3Id}
$\#\Classes^\fourdots\left(n,\big\{ \{123,132,213\} \big\}\right) = 2^{n-1}$ for $n \geq 1$.
\end{theorem}
\section{Adjacent transformations}\label{sec:adj}
As mentioned in the introduction, this section contains our most
interesting results and proofs. The first rediscovers sequence A010551
from Sloane~\cite{OEIS}.
\begin{theorem}\label{thm:P2bId}
$\#\mathrm{Eq}^\twolines\left(\iota_n,\big\{ \{123,213\} \big\}\right) = \floor{n/2}!\ceiling{n/2}!$
for $n\geq 1$.
\end{theorem}
\begin{proof}
Generically stated, our rules in this case allow the transposition of
any two adjacent elements if the element immediately to their right is
bigger than both of them. Applying these successively to $\iota_{n}$,
we note first that the largest element, $n$, never comes unglued from
the right end, because there is nothing to enable it; therefore, $n-1$
must stay somewhere in the last three positions (as only $n$ can enable
its movement). Similarly, $n-2$ remains somewhere in the last five,
$n-3$ within the last seven and so on; such restrictions apply to the
largest $\floor{n/2}$ of the elements. This limits the number of
permutations potentially reachable to $\floor{n/2}!\ceiling{n/2}!$: placing
the elements from largest to smallest, one has a choice of
$1,2,3,\ldots,\floor{n/2},\ceiling{n/2},\ldots,3,2,1$ positions to put each
element.
Next we will show that all permutations conforming to
these restrictions are indeed reachable from $\iota_{n}$. We will do this in
two stages. In Stage~1 we move each of the large, constrained
elements as far left as it can go. (In the most natural way to
achieve this, the smaller, unconstrained elements remain in
their natural increasing order, although we shall see that this
does not matter as they can then be permuted freely.) In Stage~2
we construct the target permutation two elements at a time,
working from left to right.
\textit{Stage~1: Maximally spread out the $\floor{n/2}$ largest
elements. \/} First move $\floor{n/2}+1$ one step left, using a move of
type $123 \rightarrow 213$, in which $\floor{n/2}+2$ plays the role of
the facilitating ``3''. In the same way, move the element
$\floor{n/2} + 2$ to the left, continuing until the entire block
$\floor{n/2},\dots,n-1$ has been shifted one to the left. The element
$n-1$ has now reached its leftmost permitted position, and will remain
in place as we now similarly transform the block $\floor{n/2},\dots,n-2$.
This moves $n-2$ as far left as it will go, and we now transform the next
smaller block, etc. Continue until reaching a permutation which alternates the
subsequences $1,2,\dotsc \floor{n/2}$ and $\floor{n/2}+1,\dotsc ,n$ (e.g.,
$15263748\in S_{8}$ or $516273849\in S_{9}$).
This places each constrained element (in the latter subsequence) as far
left as possible. These elements will now serve as a ``skeleton''
enabling the construction of the target permutation.
\textit{Stage~2: Construct the target permutation. \/} The key observation
making this stage possible is that
the small, unconstrained elements can be freely moved about while leaving
the large elements in the skeleton fixed. For if
$\{a,b\} < X < Y$, we can always execute the following sequence of
moves: $a\mathbf{XbY} \rightarrow \mathbf{abX}Y \rightarrow
b\mathbf{aXY} \rightarrow bXaY$.
In the case where $n$ is odd, we may consider the leftmost element in the
skeleton to be in position 3, and the two small elements in positions
1 and 2 can be interchanged if desired.
Now we examine the target permutation and move the required element(s)
into the first position (if $n$ is even), or the first two positions
(if $n$ is odd). At this point, the elements occupying the next two
positions are reclassified as small, so that the skeleton terminates
two positions further to the right, and we continue by placing and
ordering the next pair of elements. By continuing two elements at a time, we
can build the entire target permutation.
\end{proof}
\begin{eg}
To reach the target permutation $452637819$ according to the above
scheme we would apply the following moves. The numbers indicated in
\textbf{bold} are about to be transposed, either by a standard
$\mathbf{12}3\rightarrow \mathbf{21}3$ move, or by the move
(suppressing intermediate steps) $ \mathbf{a}X\mathbf{b}Y\rightarrow
\mathbf{b}X\mathbf{a}Y$ described
above.
\[
\begin{aligned}
123\mathbf{45}6789 &\rightarrow 1235\mathbf{46}789 &\rightarrow 12356\mathbf{47}89 &\rightarrow 123567\mathbf{48}9 &\rightarrow 12\mathbf{35}67849 &\rightarrow 125\mathbf{36}7849 \\
&\rightarrow 1256\mathbf{37}849 &\rightarrow 1\mathbf{25}673849 &\rightarrow 15\mathbf{26}73849 &\rightarrow \mathbf{15}6273849 &\rightarrow \cred{5}1\cred{6}2\cred{7}3\cred{8}4\cred{9} \\
51627\mathbf{3}8\mathbf{4}9 &\rightarrow 516\mathbf{2}7\mathbf{4}839 &\rightarrow 5\mathbf{1}6\mathbf{4}72839 &\rightarrow \mathbf{54}6172839 &\rightarrow 456\mathbf{1}7\mathbf{2}839\\
&\rightarrow 45\mathbf{62}71839 &\rightarrow 45267\mathbf{1}8\mathbf{3}9 &\rightarrow 4526\mathbf{73}819 &\rightarrow 452637819
\end{aligned}
\]
\end{eg}
\begin{theorem}\label{thm:P5bId}
Let $n$ be an integer $\geq 3$, and for any odd positive integer $m$ set
$m!! = 1\cdot 3\cdot \dotsb \cdot m$, the product of odd natural numbers
less than or equal to $m$. Then
\[
\#\mathrm{Eq}^\twolines\left(\iota_n,\big\{ \{123,132,321\} \big\}\right) =
\mycases{ \frac{3}{2}(k)(k+1)(2k-1)! & \kern -6.2pt for $n=2k+1$ odd. \cr
\frac{3}{2}(k)(k-\frac{1}{3})(2k-2)! - (2k-3)!! & \kern -6.2pt for $n=2k$ even.\cr}
\]
\end{theorem}
\begin{proof}
As in the previous proof, we begin by giving a set of necessary conditions
for the a permutation to be reachable from $\iota_{n}$, then show how to reach each
such permutation, thereby proving that our conditions in fact
characterise $Eq^\twolines\left(\iota_n \right)$.
The first restriction is that the element 1 must occupy
a position of odd index, because it can only participate
in a move as a ``1'', and every move either leaves it
fixed or moves it by two positions.
The second restriction is that the element 2 cannot occupy
a position of odd index to the left of 1, because if it
winds up to the left of 1, its last move there was
$132 \rightarrow 321$, and since 1 is always in a position of odd index,
this places 2 into a position of even index. Then it stays
on the left of the 1, so it must play the role of ``1'' in
any future swaps, again preserving the parity of its position.
Let us call the class of permutations thus described $\mathcal{A}_n$, the
class of \emph{admissible} permutations.
Now in the case where $n=2k+1$ is odd, this characterization is exact,
so we will first complete the proof for odd $n$. In the case where $n=2k$
is even, there are a small number of exceptional permutations which must
be excluded; we will turn to these at the end of the proof.
\textit{Case~1: $n$ is odd. \/} First we count the number of admissible permutations:
If the 1 is in position 1, then the 2 can be in any of
$n-1$ positions, and the remaining $n-2$ elements can be arranged
in $(n-2)!$ ways.
If the 1 is in position 3, then the 2 can be in any of
$n-2$ positions; if the 1 is in position 5, then in any of
$n-3$ positions, and so forth, while in each case, the remaining
$n-2$ elements can be placed freely.
Summing over the possible locations for the element 1, we arrive
at the given formula for odd $n$, and also at the formula for even $n$
upon suppression of the double-factorial correction term.
For example $\mathcal{A}_{5}$ consists of 54 permutations:
all 24 of the form $1****$, 18 of the form $**1**$ (all but the six of
the form $2*1**$), and 12 of the form $****1$ (all but those of the form
$2***1$ or $**2*1$).
It remains to show that all admissible
permutations are in fact reachable. We do this in two stages.
\textit{Stage~1:\/} First we will show that all permutations beginning with a 1 are
reachable from the identity. We proceed in steps; after each, we will have
a monotonically increasing initial segment, followed by
a segment that matches the target permutation. This segment gets
created from right to left, each step increasing the length of the
completed segment by 1 by selecting and moving one element from the
increasing segment to the left end of the completed segment.
Note that within an increasing segment, the concatenation of moves
$a\mathbf{bxy}\rightarrow \mathbf{ayx}b\rightarrow axyb$ allows a
selected element $b>1$ to move two positions rightward while maintaining
that the segment to its left is increasing. So if the target position
for $b$ is an even number of positions away, an appropriate number of
such moves will suffice. If $b$ is an odd number of positions away,
first apply the move $\mathbf{abx}y\rightarrow axby$, then proceed as
before. This shows that we can reach any permutation that starts with a
1 from $\iota_{n}$.
\textit{Stage~2:\/} To show that we can get to the identity from an
arbitrary admissible permutation, it remains to show that the element 1
can always be moved to the front of such a permutation. In fact we
only need show that the element 1 can always be moved \emph{toward} the
front (necessarily by two positions), and then we can just move it
repeatedly until it is \emph{at} the front.
If the 1 is at the very end of the permutation, the 2 must be to its
left and in a position of opposite parity. Move the 2 rightward using
moves $123\rightarrow 321$ or $132\rightarrow 321$ (the 2 functioning as
a ``1'') until it is adjacent to the 1; the 1 can then be moved
leftward.
We use the term \textit{$k$-factor} here and in later proofs as
shorthand for ``length $k$ subsequence of adjacent elements'' of a
permutation. If the 1 is not at the very end of the permutation, then
consider the 5-factor centred on the 1. (At this point, we
are relying on the assumption that $n$ is odd, because the largest
element occupies a position of odd index, and therefore if it is
not at the end of an odd permutation, it must be at least two
positions away from the end, guaranteeing the 5-factor
which we need. We will return to this point when we consider the
even case below.) There are 24 cases.
For 18 of these cases, we know that we can convert this segment
to an increasing one (or to any other permutation beginning
with a 1) using the analysis for $n=5$, which is easy to check by hand.
The cases which cannot be handled are those of the form 2*1**.
We will add a preprocessing step to make sure that we are not in
such a case. Namely, we will locate the element 2 (the actual 2)
and move into one of the spaces indicated by a $*$.
If the 2 is somewhere to the left of the 1 then the same argument used
above in the case of permutations ending in 1 again shows that the 1 can
be moved leftward.
If the 2 is somewhere to the right of the 1, we will go and fetch it as
follows. Move the 1 to the \emph{right} until it is either one or two
positions left of the 2. We do this by moving it two positions at a
time, using either $123 \rightarrow 321$ or $132 \rightarrow 321$, as
required. This leaves behind a consecutive trail of elements in which
each odd-position element is larger than the even-position element which
follows it. We will call these ``odd/even descents''.
If the 2 was in an odd position, we will arrive at $1x2$, which we
correct to $12x$. If the 2 was in an even position, we will arrive
at $12x$ directly.
Now we pull both the 1 and the 2 back through the odd/even descents by a
sequence of consecutive moves of the form $\mathbf{yx1}2\rightarrow
\mathbf{1xy}2\rightarrow 1\mathbf{yx2}\rightarrow 12yx$, where $y>x>2$.
(We may also apply $12\mathbf{yx}\rightarrow 12xy$ if we wish, but this
isn't necessary.) This brings us to a permutation in the same
equivalence class, where the 5-factor has been modified to **12*. But
we know that just working within this 5-factor, we can use the $n=5$
case to modify it to the form $12345$ (where the 1 and 2 are actual
values, the others relative values). In particular, we have moved the
actual 1 two spaces to the left. Doing this repeatedly gets us to a
permutation beginning with 1, which we have seen in Stage~1 is
equivalent to $\iota_{n}$.
\textit{Case~2: $n=2k$ is even. \/}
In the even case, we need to describe an additional class of
permutations that not reachable from $\iota_{n}$. Let $\mathcal{X}_{n}$
consist of all permutations obtainable as follows: Fill the positions in
order $n-1,n,n-3,n-2,n-5,n-4 \ldots 3,4,1,2$, according to the following
rule. When filling positions of odd index, the smallest available element
must be chosen; the subsequent selection of an element to place to its
right is then unconstrained. Thus 1 must be placed in position $n-1$,
and the element placed in position $n$ could be any other number; however, if it
is not 2, then the 2 is immediately placed in position $n-3$; otherwise
$3$ is placed in this position.
For example, $\mathcal{X}_{4}=\{3412, 2413, 2314\}$ and
\[
\mathcal{X}_{6}=\left\{
\begin{matrix}
563412 & 562413 & 562314 & 462315 & 452316 \\
463512 & 462513 & 362514 & 362415 & 352416 \\
453612 & 452613 & 352614 & 342615 & 432516 \\
\end{matrix}
\right\}\,.
\]
The number of permutations in the class $\mathcal{X}_n$ just described is $(n-1)!!$.
As we will see next, none of them is reachable. However, it is also true
that most of them are not in $\mathcal{A}_n$, and therefore have not been
included in the enumeration; this is because most of
the permutations in $\mathcal{X}_n$ have the $2$ in position $n-3$, which is a
position of odd index to the left of the $1$. The only permutations in
$\mathcal{X}_n$ which we have counted, and which therefore must be subtracted off,
are the ones where the $2$ is in position $n$, of which there are $(n-3)!!$.
To see that none of the permutations in $\mathcal{X}_n$ is reachable, consider
their $3$-factors. Every $3$-factor centred on a position of odd index
is either a $213$ or a $312$, because the middle element was placed before
either of its neighbours, and was the minimal available element at the
time it was placed. And every $3$-factor centred on a position of even
index is a $231$, because the elements in positions of odd index, which
are the minimal elements, descend from left to right. Therefore permutations
belonging to $\mathcal{X}_n$ contain \emph{no} factors of form $123$, $132$, or
$321$, and are therefore isolated by the relation, each one being a singleton
equivalence class. In particular they are not in the equivalence
class of the identity.
Now we have to consider which permutations in $\mathcal{A}_n$ are not in fact
reachable. The proof for odd $n$ almost carries through completely; indeed,
as remarked, it only fails when the element 1 lies in the penultimate
position $n-1$. We have already seen that the permutations belonging
to $\mathcal{X}_n \cap \mathcal{A}_n$ are not reachable; we will show that all
others are. Take any permutation $\pi \not\in \mathcal{X}_n$, but with the
minimal element $1$ placed in position $n-1$. Checking the conditions
from right to left, suppose the element $\pi_{j}=y$ represents
the last time that we were in compliance with the conditions, and suppose
$\pi_{i}=x$ is the first minimal element which has not
gone where it should go. That is, all odd positions from $j$ to $n-1$ are
occupied by elements which are left-to-right minima, but the smallest
element situtated in positions $1$ through $j-1$ is not in position
$j-2$, as expected, but in position $i$ with value $x$.
As before, all we need to do is show that we can move the element $1$
to the left. This exploits two facts: that $x$ is the minimal element in
a lefthand region, and the righthand region is alternating.
Because the righthand portion of $\pi$, from position $j$ onwards, is
alternating, with every step from an odd to an even position being an
ascent, and every step from even to odd being a descent, we will have
a particular interest in a certain type of $3$-factor beginning in
a position of odd index. Namely, we will refer to a $3$-factor
$\pi_{h},\pi_{h+1},\pi_{h+2}$ as an \emph{odd 321} if
$\pi_{h}>\pi_{h+1}>\pi_{h+2}$ and if $h$ is odd. Note that an odd $321$
beginning in position $n-3$ is exactly what we need, because either option
for replacing it shifts the element $1$ from position $n-1$.
First, take the element $x$ and use moves $\rightarrow 321$ to shift
it rightward, two positions at
a time, until it arrives in position $j-2$ or $j-1$. This is possible
because $x$ is moving through a region in which it is itself the minimal
element.
Now $j-2$ is an odd position, so if $x$ has reached position $j-2$
then positions $j-4,j-3,j-2$ now form an odd $321$. Alternatively, if $x$
has reached position $j-1$ then positions $j-2,j-1,j$ now form an odd $321$,
because the second and third of these positions are occupied by $x$ and $y$
and $y<x$. We will show that we can propagate either of these odd $321$s
rightward until they capture the smallest element, which can then be moved.
In either case, we have an odd $321$, followed, two positions later, by
an element which is smaller than everything to its left. This gives us,
in other words, a configuration $432-1$, which, filling in the blank,
might actually be (a) a $54231$, (b) a $53241$, or (c) a $43251$. Check that the
following moves are available in each case:
(a) $54231 \rightarrow 24531$; (b) $53241 \rightarrow 23541$;
(c) $43251 \rightarrow 23451 \rightarrow 25431$.
Note that these moves each replace a configuration which begins with
an odd $321$ by one which \emph{ends} with an odd $321$. And, because
of the placement of the left-to-right minima, this new odd $321$ either
terminates with the smallest element $1$, or again has another left-to-right
minimum two positions to its right.
Therefore we can propagate the $321$ rightward until it reaches the smallest
element; therefore we can move the smallest element; therefore the
permutations not belonging to $\mathcal{X}_n$ are in fact reachable.
This completes the missing step in the proof for even $n$.
\end{proof}
\begin{theorem}\label{propo:P4bId}
$\#\mathrm{Eq}^\twolines\left(\iota_n,\big\{ \{123,321\} \big\}\right) = \dbinom{n-1}{\lfloor (n-1)/2 \rfloor}$.
\end{theorem}
\begin{proof}
We claim that the permutations in this class are direct sums of singletons
and of blocks of odd size greater than one, where within each
block the even elements (with respect to the block) are on the
diagonal, and the odd elements form an \textit{indecomposable}
$321$-avoiding permutation.
Let us call the set that we have just described $\mathcal{A}_n$.
Because all the even elements within a block are fixed points of the
permutation, the indecomposability of the odd elements is equivalent to
the indecomposability of the entire block.
First we will show that $\mathcal{A}_n$ is closed under $123
\leftrightarrow 321$; since the identity is in $\mathcal{A}_n$ this will
establish that the equivalence class of the identity is a subset of
$\mathcal{A}_n$. Then we will show that we can return to the identity
from any permutation in $\mathcal{A}_n$, which will establish that the
two sets are identical. Finally we will use generating functions to
enumerate $\mathcal{A}_n$.
Let $\pi$ be an arbitrary permutation belonging to $\mathcal{A}_n$. By
definition, $\pi$ is a direct sum of singleton blocks and of larger
blocks having a specific form. We will call any non-singleton block of
$\pi$ \emph{large}. Unless $\pi$ is the identity, it contains at least
one large block. Note that large blocks always begin with descents: for
if the first element of the block were on the diagonal, we could split
the block immediately after it to obtain a direct sum decomposition;
therefore, the first element is below the diagonal (i.e., is an
excedance) but the second element is on the diagonal. For symmetric
reasons, large blocks end with descents as well.
First we show that any application of $123 \rightarrow 321$ to $\pi$
produces an element $\pi '$ in $\mathcal{A}_n$. Consider the different ways that a 3-factor
$\pi_i,\pi_{i+1},\pi_{i+2}$ of form $123$ might occur within $\pi$.
\textit{Case~(a)} All three elements are in singleton blocks; then the result is the
unique large block of size 3 permissible within elements of
$\mathcal{A}_{n}$, namely
$\begin{pmatrix}
0&0&1\cr 0&1&0\cr 1&0&0\cr
\end{pmatrix} $.
\textit{Case (b)} Exactly two of the elements (necessarily the first two
or the last two) are in singleton blocks. Assume without loss of
generality that it is the last
two, i.e., $\pi_{i+1}=i+1$ and $\pi_{i+2}=i+2$, and that $\pi_i < i$
belongs to a large block $B$ of size $2k+1$. Since $\pi_{i}$ is the last
element of $B$, it must be an odd element within the
block. The replacement produces a larger block $B'$ of size $2k+3$.
The $k$ even elements of $B$, along with $\pi_{i+1}$ are diagonal
elements that remain unchanged by the transformation, so all the even
elements of $B'$ lie on the diagonal. The block $B'$ must be
indecomposable, because any breakpoint before $\pi'_{i}$ would already
have been a breakpoint for $B$ itself, and no breakpoint can occur
thereafter since $\pi'_{i}>\pi'_{i+1}>\pi'_{i+2}$.
Finally, $\pi'_{i}$ is the largest element of the block, so could only
play the role of ``3'' in a 321 pattern, but there is only one odd
element to its right in the block. So any 321 pattern of odd elements
in $B'$ that did not already exist in $B$ must use $\pi'_{i+2}$ as ``1''
and odd elements to the left of $\pi'_{i}$ for ``3'' and ``2''. But
then these elements (which haven't moved) together with $\pi_{i}$ would
have formed a 321 in $B$, which wasn't allowed.
\textit{Case~(c)} Just one element is in a singleton block.
This can't be the third (or, symmetrically, the first) element, because
if $\pi_i$ and $\pi_{i+1}$ are the final two elements of a large
block, then $\pi_i > \pi_{i+1}$, so our 3-factor is not a 123. So it must be
$\pi_{i+1}$ which is the singleton, while the other two elements belong
to two large blocks. The replacement merges these three blocks into one;
the even elements, including $\pi_{i+1}$, remain on the diagonal, and as
in the previous case any point at which the new block split would also
imply a decomposition of one of the old blocks at the same
position. The odd elements of $\pi '$ are $321$-avoiding because if a $321$
contained just one of $\pi'_i$ or $\pi'_{i+2}$ then it would be
pre-existing (with $\pi_{i+2}$ or $\pi_{i}$ respectively). If it
contained both, then the third element in the pattern would be either on
the left and too large for the old lefthand block, or on the right and
too small for the old righthand block.
\textit{Case (d)} The three elements are all within a single large block
$\mathcal{B}$.
First we claim that the middle element, $\pi_{i+1}$ must be in an even
position (within $\mathcal{B}$). Otherwise, $\pi_{i}$ and $\pi_{i+2}$
would be in even positions, hence on the diagonal, and the 123 form of
the 3-factor would mean $\pi_{i+1}$ was also on the diagonal; thus,
$\mathcal{B}$ would have to be of size at least 5. Now if all elements to
the left of $\pi_{i}$ were smaller than it, $\mathcal{B}$ would split into
summands before position $i$. But if some $\pi_{j}$ is greater than
$\pi_{i}$ (for some odd $j<i$) it
must be greater than $\pi_{i+2}$, forcing a compensatory $\pi_{k} < \pi_{i}$ for
some odd $k>i+2$. But then $\pi_{j}$, $\pi_{i+1}$, $\pi_{k}$ formed a
321-pattern of odd positions within the block, contrary to hypothesis.
The claim follows.
Now the replacement $\pi_{i}\pi_{i+1}\pi_{i+2}\rightarrow
\pi_{i+2}\pi_{i+1}\pi_{i}$ cannot create a new direct sum decomposition
since it is increasing the left element and decreasing the right one.
Suppose that somehow this move created a $321$ among the odd elements
(within $\mathcal{B}$). If it only used one of $\pi_{i}, \pi_{i+2}$,
then it must have been pre-existing with the other one, contrary to
hypothesis. If it used both, then without loss of generality assume
$\mathcal{B}$ contains an element $x$ to the left of the replaced
3-factor, but larger than $\pi_{i+2}$. Because $x$ is also greater than
$\pi_{i+1}$, it uses up one of the odd values greater than the diagonal
element $\pi_{i+1}$, meaning that there must be a $y$ to the right of
$\pi_{i+1}$ but smaller than it, and then $x,\pi_{i+2},y$ is a
pre-existing $321$. The case where $\mathcal{B}$ contains an element
$y$ to the right of the replaced 3-factor, but smaller than $\pi_{i}$
follows by symmetry.
\textit{Non-Case (e)} The last possibility to consider is that the
3-factor is split across two adjacent large blocks, necessarily with two
elements at the start or end of one of the large blocks. But this is
ruled out because large blocks begin and end with descents.
Note that in each of these cases the replacement $123
\rightarrow 321$ winds up gluing together all the blocks of $\pi$ which
it straddles, leaving the same number or fewer blocks in $\pi'$. In
particular, the replacement may glue together blocks, but never splits
any apart.
Now consider applications of $321 \rightarrow 123$ within a permutation
$\rho \in \mathcal{A}_n$ to obtain a new permutation $\hat{\rho}$. Clearly,
any adjacent $321$ must lie within a single
block, as in any two blocks, all the elements in the block to the right
are larger than all the elements in the block to the left. Because the
even elements within a block increase monotonically, the $321$ is
composed of odd, even, odd elements. An analysis similar to that given
above shows that any such
transformation is simply the reverse of one of the cases (a--d)
described above, so $\hat{\rho}$ is always in $\mathcal{A}_{n}$.
Now we need to show that we can use these transformations to return to
the identity from any permutation $\sigma$ in $\mathcal{A}_n$. We first
claim that every large block of $\sigma$ contains a $321$ as a factor.
For the first element of the block must lie below the diagonal and the
last element must lie above it; therefore, there are two consecutive odd
elements in the block with the first below and the second above the
diagonal. Together with the even element which separates them, and which
lies \emph{on} the diagonal, this forms a $321$.
Unless $\rho $ is itself the identity, it contains a large block, and
therefore a $321$. Replacing this $321$ with a $123$ yields a
permutation $\hat{\rho}$ having strictly fewer inversions than $\rho$. But
as $\mathcal{A}_n$ is closed under such replacements, we know that $\hat{\rho}$
also belongs to $\mathcal{A}_n$, and therefore is either the identity or
else contains a $321$. By iterating this process, we must eventually
arrive at a permutation having no inversions, namely the identity.
This establishes that $\mathrm{Eq}^\twolines\left(\iota_n,\big\{ \{123,321\}
\big\}\right)=\mathcal{A}_{n}$, so all that is left is the enumeration
of these classes. It is an easy exercise
\cite[(n$^{6}$)]{StanleyCatAdd} or \cite[p.~15]{ClaKit} that the number of \emph{indecomposable}
$321$-avoiding permutations on $m+1$ elements is the Catalan number
$c_{m}={\frac{1}{m+1}}{{2m}\choose{m}}$. This is also the number of possible
blocks of size $2m+1$.
We define the following three generating functions, which enumerate
central binomial coefficients of even order (E), of odd order (O),
and the Catalan numbers (C).
\begin{eqnarray*}
E(x) & = \frac{1}{\sqrt{1-4x}} & = 1 + 2x + 6x^2 + 20x^3 + 70x^4 + \dots \\
O(x) & = \frac{\frac{1}{\sqrt{1-4x}}-1}{2x} & = 1 + 3x + 10x^2 + 35x^3 + 126x^4 + \dots \\
C(x) & = \frac{1-\sqrt{1-4x}}{2x} & = 1 + x + 2x^2 + 5x^3 + 14x^4 + \dots
\end{eqnarray*}
The statement of the theorem is equivalent
to showing that $E(x)=\sum_{n\geq 0} A_{2n+1}x^{n}$ and $O(x)=\sum_{n\geq
0} A_{2n+2}x^{n}$, where we set $A_{n}:=\#\mathcal{A}_{n}$.
Now a reachable permutation of even size $2n+2$ is the direct sum of
an indecomposable block of size $2i+1$ ($i\geq 0$) and a reachable
permutation of odd size $2(n-i)+1$. This translates into the
recursion/convolution
\[
A_{2n+2}=\sum_{i=0}^{n}c_{k}A_{2(n-i)+1}
\]
which is equivalent to $O(x)=E(x)C(x)$, and which is also easily verified from
the closed-form expressions for these generating functions. Similarly,
a reachable permutation of odd-size $2n+1 $ is
the direct sum of an indecomposable block of size $2i+1$ and a
reachable permutation of even size $2(k-i)$, corresponding to the
easily-verified equality of generating functions $E(x)=(1+xO(x))C(x)$.
This completes the proof.
\end{proof}
Although the above proof seems natural enough from the structure of the
equivalence class $\mathcal{A}_{n}$, the simple form of the enumeration
as a single binomial coefficient begs the question of whether there is a
more direct (perhaps bijective) argument.
The next theorem provides independent proofs of two results which
appeared 10 years ago in [CEHKN].
\begin{theorem}\label{propo:P3bClassInv}
(a) $\#\Classes^\twolines\left(n,\big\{ \{123,132,213\} \big\}\right) = \mathrm{inv}_{n}$, the
number of involutions of order $n$.
(b) $\#\mathrm{Eq}^\twolines\left(\pi,\big\{ \{123,132,213\} \big\}\right)$ is odd for all $n$ and for each $\pi \in S_n$.
\end{theorem}
\begin{proof}
Write each involution in $\tau \in \mathrm{Inv}_{n}\subseteq S_{n}$ canonically
as a product of 1-cycles and 2-cycles, with the elements increasing within each 2-cycle,
and with the cycles in decreasing order of largest element. Omitting
the parentheses, we view the resulting word $D(\tau)$ as a
permutation. Let $\mathcal{D}_{n}:=D(\mathrm{Inv}_{n})$ be the image of this map
(which is easily reversible by placing parentheses around the ascents of
$\sigma \in \mathcal{D}_{n}$). We claim that this is a canonical set of
representatives for the equivalence classes of $S_{n}$ under $P_{3}= \{
\{123,132,213\} \}$ transformations.
Each permutation $\pi\in S_{n}$ can be transformed to an element of
$\mathcal{D}_{n}$ as follows:
if $n$ is at the front of $\pi$, it must stay there. (This corresponds
to having $n$ as a fixed point of the involution.) Otherwise, use $123
\rightarrow 132$ and $213 \rightarrow 132$ (at least one of which is
possible at each step) to push $n$ leftward into
position 2, which is as far as it will go. The element which is thus
pushed into position 1 is the minimal element $m$ which was to the left of
$n$ to begin with. This is because $m$ can never trade
places with $n$ under the given operations, as 1 is left of 3 in all of
123, 132 and 213. Leaving the leftmost 1-factor $n$ or 2-factor $mn$
fixed, proceed inductively among the remaining elements, at each step
moving the maximal remaining element as far left as possible. The end
result of this deterministic procedure is a permutation $L(\pi)\in
\mathcal{D}_{n}$. This shows that the number of $P_{3}$-equivalence
classes is at most $\mathrm{inv}_{n}=\#\mathcal{D}_{n}$.
To show that they are the same, it remains to show that each $\pi$ can
be transformed to a \emph{unique} member of $\mathcal{D}_{n}$, or
equivalently that it is not possible to move from one member of
$\mathcal{D}_{n}$ to another using $P_{3}$-moves.
We will prove this by induction on $n$. At the same time we will prove
statement (b) of the theorem. Assume as an induction hypothesis that
both statements have been demonstrated for $n-1$ and $n-2$. It is
straightforward to check the base cases by hand. For $n=3$ the four
equivalence classes are $P_{3}$ and three singleton classes. For $n=4$
the classes are $\{1234,1243,1324,2134,\mathbf{1423},1342,2143,3142,2314
\}$, $\{1342, 3124, \mathbf{1432}, 3142, 3214 \}$, $\{4123,
\mathbf{4132}, 4213 \}$, $\{2341, \mathbf{2431}, 3241\}$, and six
singletons: $\mathbf{\{ 2413\}, \{3412\},}$ $\mathbf{\{4312\},}$
$\mathbf{4231}$, $\mathbf{3421}$, $\mathbf{4321}$. (The
elements in \textbf{bold} are the class representatives within
$\mathcal{D}_{n}$.)
First note that if the largest element, $n$, is at the front of a
permutation, then it is immobile under $P_{3}$-moves. Thus the
equivalence classes split into two kinds: \emph{special} equivalence
classes, in which $n$ is at the front of each permutation in the class,
and \emph{ordinary} equivalence classes, in which $n$ is never at the
front. Moreover it is obvious that the special equivalence classes for
$S_n$ correspond exactly to \emph{all} the equivalence classes for
$S_{n-1}$ upon deletion of the first elements; therefore,
the truth of both (a) and (b) as they apply to the
\emph{special} equivalence classes follows by induction.
Next we will look at the \emph{ordinary} equivalence classes. For
convenience of exposition, consider a (directed) graph in which the
vertices correspond to the permutations in $S_n$, and there is a blue
(directed) edge from $\pi$ to $\pi'$ if $\pi'$ can be obtained from
$\pi$ by applying $123 \rightarrow 132$, and similarly a red edge for
each $213 \rightarrow 132$, and a green edge for each $123 \rightarrow
213$. A blue edge just corresponds to a green edge followed by a red
one, and indeed the edges always appear in matched sets: the appearance
of a $213$ in a permutation implies an incident green edge pointing in
and a red edge pointing out, and also a blue edge making the chord of
this triangle (and similarly for appearances of $123$ and $132$). The
equivalence classes in which we are interested are the (undirected)
connected components of this graph.
Now consider the forest of rooted trees which one obtains by taking only
those red and blue edges in which the element $n$ plays the role of the
``3''. The roots (i.e., sinks) of these trees are exactly the
permutations in which the $n$ has moved to position $2$, which is as
far left as it will go within an ordinary equivalence class. More generally,
if $\pi_{k}=n$, then $\pi$ lies on level
$k-2$ of the tree. (We can say that it has \emph{energy}
$E(\pi)=k-2=\pi^{-1}(n)-2$.) Note that blue and red edges reduce the
energy by one, while green edges leave it unchanged.
Each vertex in this forest has either zero or two children, because if it
has a blue child (obtained by travelling backwards along a blue edge)
then it also has a red child, and vice versa.
Each permutation $\pi$ lies on a unique directed path to the
root of its tree, which we will call the \emph{ground state} of $\pi$,
$g(\pi)$. Note that $g(\pi)_{2}=n$, while $g(\pi)_{1}$ is the smallest
element $m$ to the left of $n$ in $\pi$.
Because each node has either zero or two children, each rooted tree has
an odd number of nodes; indeed all of its level-sums are even except the
zeroth level sum, which corresponds to the root vertex (i.e., ground
state).
Now we will create larger classes as follows: declare two ground states
$\tau$ and $\sigma$ \emph{similar} if $\tau_{1}=\sigma_{1}$ and
$\tau_{3}\dotsb \tau_{n}$ is $P_{3}$-equivalent to $\sigma_{3}\dotsb
\sigma_{n}$ regarded (in the obvious way) as members of $S_{n-2}$. For
$m\in [n-1]$ and $\nu \in \mathcal{D}_{n-2}$, let
$\mathcal{K}(m,\nu)$ be the (disjoint!) union of all trees with similar ground
states $\tau $, where $\tau_{1}=m$ and $\tau_{3}\dotsb \tau_{n}$ is
$P_{3}$-equivalent to $\nu$. Note that this gives us a total of
$(n-1)\mathrm{inv}_{n-2}$ equivalence classes, in agreement with the well-known
recursion: $\mathrm{inv}_{n}=\mathrm{inv}_{n-1}+(n-1)\mathrm{inv}_{n-2}$. (The special
equivalence classes account for the first summand.)
We claim that these larger classes $\mathcal{K}(m,\nu)$ are exactly (the
vertex sets of) the connected components of our directed graph; that is,
there are no directed edges in the graph which escape from one class to
another. Once this is shown, then by induction there is a unique member
of the class $\mathcal{D}_{n-2}$ of canonical permutations among the
ground states in a large component, to which we prepend $mn$ to obtain
the unique representative of $\mathcal{K}(m,\nu)$ within
$\mathcal{D}_{n}$.
Furthermore, each $\mathcal{K}(m,\nu)$ will then be of odd size, because
each rooted tree has odd size, having all level-sums even except the one
corresponding to the ground states, and because the number of rooted
trees in the union is odd by the induction hypothesis for $n-2$.
So suppose there is an edge (of any colour) from a $\pi \in
\mathcal{K}(m,\nu)$ to $\pi'\in \mathcal{K}(m',\nu')$.
Since this move does not involve moving the largest element $n$,
$\pi$ and $\pi'$ have the same energy. Our goal is to show that
$m=m'$ and $\nu $ is $P_{3}$-equivalent to $\nu'$. The former follows
from our earlier description of $m$ as the minimum element lying to the
left of $n$ in $\pi$, because $\pi$ and $\pi'$ have the same set of
elements to the left of $n$. The latter requires an analysis of the
cases that can arise as $\pi$ and $\pi'$ move towards their ground
states in their respective trees.
As the $n$ moves leftward through each of the two permutations
(following red and/or blue edges toward their respective ground states)
then it sometimes encounters identical elements and therefore has the
same effect; eventually it encounters the positions where the
difference lies, having swept before it the
minimal intervening element, $b$. What happens from this point forward
depends on how $\pi$ and $\pi'$ differ, and the relative value of $b$.
To clarify the cases, let the three values where the difference was
applied be $d < f < h$, and designate $b$ by one of $C,E,G$ or $I$
(where $C<d<E<f<G<h<I$), depending on its relative order within the
factor. For example, at some point along the path
from $\pi$ to $g(\pi)$ we may see a permutation containing the factor
$dfhCn$, while at the same energy level on the path from $\pi'$ to
$g(\pi')$ we see instead $dhfCn$ (having followed a blue edge), the two
permutations being otherwise identical. Advancing the element $n$ three further
steps to the left, we arrive in the first instance at $Cndfh$ and in the second
instance at $Cndhf$; the $n$ then continues forward all the way to
position $2$ (zero energy), making identical moves in each case. The
resulting ground states $g(\pi)$ and $g(\pi')$ differ only by a (blue)
move $dfh \rightarrow dhf$, so $\nu =\nu '$.
Here is a table of the cases that arise given the four possible relative
values of $b$; blue edges are the composition of green with red.
\begin{eqnarray*}
Input: dfhCn \hspace{5mm}{\color{green}\rightarrow} & fdhCn & {\color{red}\rightarrow}\hspace{5mm} dhfCn \\
Output: Cndfh \hspace{5mm}{\color{green}\rightarrow} & Cnfdh & {\color{red}\rightarrow}\hspace{5mm} Cndhf \\
\\
Input: dfhEn \hspace{5mm}{\color{green}\rightarrow} & fdhEn & {\color{red}\rightarrow}\hspace{5mm} dhfEn \\
Output: dnEfh \hspace{5mm}{\color{green}\rightarrow} & dnfEh & {\color{red}\rightarrow}\hspace{5mm} dnEhf \\
\\
Input: dfhGn \hspace{5mm}{\color{green}\rightarrow} & fdhGn & {\color{red}\rightarrow}\hspace{5mm} dhfGn \\
Output: dnfGh \hspace{5mm}{=} & dnfGh & {\color{blue}\rightarrow}\hspace{5mm} dnfhG \\
\\
Input: dfhIn \hspace{5mm}{\color{green}\rightarrow} & fdhIn& {\color{red}\rightarrow}\hspace{5mm} dhfIn \\
Output: dnfhI \hspace{5mm}{=} & dnfhI & {=}\hspace{5mm} dnfhI \\
\end{eqnarray*}
Examining this table shows that the classes $\mathcal{K}(m,\nu)$
containing $\pi$ and $\mathcal{K}(m',\nu')$ containing $\pi$ and $\pi'$
have $\nu $ $P_{3}$-equivalent to $\nu'$, which completes the proof.
\end{proof}
This result is particularly striking because the
equivalence relation has the same number of classes as Knuth
equivalence, yet the two relations are materially different. For
example, for $n=3$, the equivalence classes for $P_{K}$ have sizes
1,1,2,2, whereas for $P_{3}=\big\{ \{123,132,213\} \big\}$ the sizes are
1,1,1,3. In fact the authors in \cite{Chinese} show that the corresponding
monoids (plactic and Chinese) share
the same graded Hilbert series, and they obtain a partial recurrence for the
numbers $\#\mathrm{Eq}^{\twolines}(\iota_{n},P_{3})$.
\begin{propo}[\cite{Chinese}, Cor. 4.3]\label{propo:chinese} For $n$ odd,
$\#\mathrm{Eq}^{\twolines}(\iota_{n},P_{3}) = n\cdot
\#\mathrm{Eq}^{\twolines}(\iota_{n-1},P_{3})$.
\end{propo}
The recurrence for $n$ even appears still to be open.
\section{Doubly adjacent transformations}\label{sec:dbl}
For completeness we include a brief treatment of the situation where
both indices and values are simultaneously constrained to be adjacent.
In this highly constrained situation, the permutations reachable from
the identity are easy to classify and enumerate in all cases. Since all
the treatments are similar, we can wrap them up in one proposition.
As in the previous section, we have as yet no results related to the
enumeration of equivalence classes.
The statement of this proposition makes use of the \textit{Iverson bracket}\/: [S] is equal
to 1 if the statement S is true, and 0 otherwise.
\begin{propo}\label{propo:P*Id}
$\#\mathrm{Eq}^\square(\iota_n,P_1)$ obeys the recurrence $a(n) = a(n-1) + a(n-2)$
with $a_1 = a_2 = 1$. (Fibonacci numbers $F(n)$, \cite[A000045]{OEIS}).
$\#\mathrm{Eq}^\square(\iota_n,P_4)$ obeys the recurrence $a(n) = a(n-1) + a(n-3)$
with $a_0 = 0$, $a_1 = a_2 = 1$ (\cite[A000930]{OEIS}).
$\#\mathrm{Eq}^\square(\iota_n,P_3) = F(n+1) - [n$ is even$]$.
$\#\mathrm{Eq}^\square(\iota_n,P_5)$ obeys the recurrence $a(n) = a(n-1) + a(n-2) + a(n-3)$
with $a(0)=a(1)=a(2)=1$ (Tribonacci numbers, \cite[A000213]{OEIS}).
$\#\mathrm{Eq}^\square(\iota_n,P_7) = T(n+2) - [n$ is even$]$,
where $T(n)$ obeys the recurrence
$T(n) = T(n-1) + T(n-2) + T(n-3)$ with $T(0)=T(1)=0,
T(2)=1$. (Tribonacci numbers (with different initial conditions),
\cite[A000073]{OEIS}).
\end{propo}
\begin{proof}
We begin by characterizing the various equivalence classes of $\iota_{n}$. In each
case, no element can move any further from its starting position then it
could via a single move. The resulting classes are subsets of those
layered permutations which are direct sums of anti-identities of
dimensions either 1, 2 or 3, as follows:
$P_1$ ($123 \leftrightarrow 132$): any direct sum of $\rho_1$ and
$\rho_2$ beginning with $\rho_{1}$;
$P_4$ ($123 \leftrightarrow 321$): any direct sum of $\rho_1$ and
$\rho_3$;
$P_3$ ($123 \leftrightarrow 132 \leftrightarrow 213$): any direct sum of
$\rho_1$ and $\rho_2$ including at least one $\rho_1$;
$P_5$ ($123 \leftrightarrow 132 \leftrightarrow 321$): any direct sum of
$\rho_1$, $\rho_2$ and $\rho_3$ not beginning with $\rho_2$;
$P_7$ ($123 \leftrightarrow 132 \leftrightarrow 213 \leftrightarrow
321$): any direct sum of $\rho_1$, $\rho_2$, $\rho_3$ with at least one
of odd dimension;
In each case it is easy to see that the given class remains closed under
application of the appropriate operations. It is also easy in general to
see how to reach a given target permutation from $\iota_{n}$, especially
if we cast the block sizes in the language of regular expressions. The
notation $\{xy\}$ means a single block of size either $x$ or $y$. An
asterisk following a number means zero or more copies of that number. An
asterisk following a string within $[\quad ]$ (not to be confused with
the Iverson brackets in the statement of the proposition) indicates zero
or more copies of that string.
$P_1$: The block sizes are $1\{12\}*=[12*]*$. Build each string of
blocks of the form $12*$ from right to left.
$P_4$: The block sizes are $\{13\}*$; build each block freely.
$P_3$: From any non-identity permutation with block sizes as described,
there is at least one instance of $21$ or $12$, which can be transformed
by one of the rules into a $111$. The resulting permutation has one
fewer $\rho_{2}$. Proceed inductively to transform all the $\rho_{2}$'s
(i.e., 2-blocks) to consecutive 1-blocks until the identity is reached.
$P_5$: The block sizes are $[\{13\}2*]*$. First use $123 \rightarrow 132$ to build
all the 2-blocks from right to left. Then use $123 \rightarrow 321$
to place the 3-blocks.
$P_7$: Build the 2-blocks first, as in the case of $P_3$, and then place
the 3-blocks.
One now verifies all the necessary base cases, as trivially $a_1=1$, $a_2=1$,
and $a_3=$ the size of the non-singleton block of $P_{j}$.
As for the recurrences, for $n>3$:
$P_1$: $a_n = a_{n-1} + a_{n-2}$, by appending respectively a $\rho_1$ or a $\rho_2$.
$P_4$: $a_n = a_{n-1} + a_{n-3}$, by appending respectively a $\rho_1$ or a $\rho_3$.
$P_5$: $a_n = a_{n-1} + a_{n-2} + a_{n-3}$, by appending $\rho_1$, $\rho_2$ or $\rho_3$.
$P_3$: Count all direct sums of $\rho_1$ and $\rho_2$
(obviously Fibonacci) and then subtract 1 from the even terms to remove
the special case $2*$ (all blocks of size 2).
$P_7$: Count all direct sums of $\rho_1$, $\rho_2$, $\rho_3$ to get
tribonacci numbers [A000073],
and subtract 1 from the even terms because block structure $2*$ is
disallowed. Alternatively,
verify the recurrence $a_n = a_{n-2} + U_n$, where $U_n$ is the
$P_{5}$-recurrence [A000213], by
noting that a permutation in $\mathrm{Eq}^\square(\iota_n,P_7)$ is either a $\rho_2$
prepended to a permutation in $\mathrm{Eq}^\square(\iota_{n-2},P_7)$, or else
belongs to $\mathrm{Eq}^\square(\iota_{n},P_5)$.
\end{proof}
\section{Final Remarks \& Open Questions}\label{sec:open}
Our results in this paper are just a tractable subset of questions that
could be explored within these families of equivalence relations. We
created the framework to easily allow for a number of extensions. The
connections with familiar integer sequences, pattern-avoidance in
permutations, and important combinatorial bijections indicate the value
of further work. Possible directions for further study include:
\begin{enumerate}
\item Study the sizes (and characterise if possible) all equivalence
classes $\mathrm{Eq}^*(\pi,P)$, not just for the case $\pi
=\iota_{n}$. Corresponding to each equivalence relation is the multiset
of sizes of the equivalences classes, perhaps best considered as an
integer partition of $n!$. Is the study of these of interest?
\item Allow for more generality among the (set) partition $P$ of $S_{3}$
which defines our relations. The authors in~\cite{PRW11} allow
substitution of patterns in $S_{3}$ where no element is fixed, but still
restrict to partitions $P$ consisting of exactly one non-singleton block
containing the identity $123$. Although it seems unwieldy to work with
all $B(6)=203$ possible partitions of $S_{4}$, perhaps a different
restriction that forces greater symmetry among the relations would be
useful. For example, the Knuth relations $P^{\twolines}_{K}=\big\{ \{
213,231\}, \{132,312\}\big\}$ are closed under reversal and
complementation.
\item Consider relations generated by partitions $P$ of $S_{k}$ for
$k>3$. Here one definitely needs some conditions to restrict focus to
relations of particular interest, since the Bell number $B(4!)$ is
already far too large to handle all cases.
\item Study in greater detail the structure of the graphs defined by
these relations. What can one say about their degree sequences or
diameters? How many moves are necessary in order to transform a given
$\pi $ to the identity?
\end{enumerate}
\nocite{*}
|
\section{Introduction}
Resonant phenomena have received much attention in atomic and nuclear physics and more recently in chaotic and disordered systems \cite{disordered, GG:00,Casati:99, KS:06,KS:08, JF:09, chaos}. Complex energies, $\tilde{{E}}_{\alpha}=E_{\alpha}-\frac{i}{2}\,\Gamma_{\alpha}$, which correspond to poles of the scattering matrix on the unphysical sheet, characterize resonances \cite{LandauL}. Resonances correspond to the long-lived quasi-stationary states which eventually decay to continuum while distribution of resonance widths, $P(\Gamma)$, determines decay of the corresponding survival probability with time.
In recent years, $P(\Gamma)$ has been a subject of investigations \cite{disordered,GG:00} for a, simple but much studied, discrete tight-binding one dimensional random chain which is coupled to a perfect lead at one side. A numerical study \cite{GG:00} shows that in a broad range of $\Gamma$, $P(\Gamma)\sim \Gamma^{-\gamma}$, where the exponent $\gamma$ is very close to $1$. Intuitively the $1/\Gamma$ behaviour can be deduced by assuming a uniform distribution for the localization centers of exponentially localized states \cite{Casati:99}. However, from analytic point of view one usually considers an infinitely long chain in which case the average density of resonances (DOR) has a well defined limit. For a finite size system, the difference between the DOR and $P(\Gamma)$ is the normalization by the system size \cite{KS:06,KS:08}. Recently, Kunz and Shapiro have derived analytic expression of the DOR for a semi-infinite disordered chain \cite{KS:08}. They have obtained an exact integral representation of the DOR which is valid for arbitrary lead-chain coupling strength. This has been further simplified for small lead-chain coupling strength where a universal scaling formula is found. In this limit they have proved the $1/\Gamma$-behavior of the DOR \cite{KS:06, KS:08}. Besides, for the continuous limit of this model an integral representation of DOR has been obtained \cite{JF:09}.
Kunz and Shapiro's work has established a universal $1/\Gamma$ law for arbitrary strength of disorder in a semi-infinite chain. Numerically one can verify $1/\Gamma$ law of the DOR, similar to what has been done by Terraneo and Guarnery \cite{GG:00} in finite samples for $P(\Gamma)$. Such verifications require the localization lengths to be much smaller than the size of the sample. In case of weak disorder an analytic result for the localization lengths is particularly useful. It comes from a second order perturbation theory. It states that the localization length is maximum near the middle of the energy band and is proportional to $W^{-2}$ where $W$ is the width of the disorder \cite{thou,krammac,KWegner:1981}. On the other hand, this result also leads to an interesting limiting situation where the localization lengths are much longer than the sample size. This is what we refer to as a weak disorder limit in this paper. This limit has scarcely been studied hitherto although it is relevant in the study of localization through resonances. Besides, there has been a believe for some sort of universality in the weak disorder limit. In this paper we address to this limit and derive analytic results which describe the statistics of resonances. Our work probes a fresh area and studies a weak disorder limit which has never been addressed before.
For open systems, instead of studying the scattering matrix in a complex plane we follow an alternative approach where one solves the Schr\"{o}dinger equation by describing a particle ejected from the system or equivalently with a boundary condition of outgoing waves (Siegert boundary condition \cite{Siegert}). In this approach one naturally turns up to a problem of solving a non-Hermitian effective Hamiltonian which admits complex eigenvalues $\tilde{{E}}_{\alpha}$ \cite{GG:00, KS:06, KS:08, HKF:09}. For details of such non-Hermitian effective Hamiltonians, we refer to a recent study \cite{JF:10} and references therein.
We derive the statistics which describe scattered complex energies of disordered chain around those regular ones which correspond to an open chain without any disorder (clean chain). For instance, we derive average of square of the absolute values of the shifts in complex energies from the regular ones over all realizations of the set of random site energies. Similarly we obtain results for the statistics of real and imaginary parts of those shifts. These results lead to compact expressions for long chains. To show the generality of our approach we also derive these results for the so-called parametric resonances which have been particularly useful in numerical studies \cite{GG:00}. Finally, we give numerical verifications of our analytic results.
The paper is organized as follows. Although the system and its effective Hamiltonian have been nicely explained earlier in \cite{GG:00,KS:06,KS:08}, for the sake of completeness of this paper we will describe these briefly in section II. In the same section we will also describe the exact and the parametric resonances. In Sec. III we will derive result for resonances in an open-clean chain of finite length, in terms of a polynomial equation. For long chains, we will solve this polynomial equation in the leading order of the inverse of the length. In Sec. IV we will use the perturbation theory to obtain the first and the second order corrections in the complex energies for a weak disorder. In Sec. V we will calculate statistics of the scattered complex energies. In the same section we will simplify our results for long chains and obtain compact expressions. In Sec. VI we will briefly discuss about the numerical methods to calculate complex energies of non-Hermitian effective Hamiltonians and numerically verify our analytical results. This will be followed by the conclusion in Sec. VII.
\begin{figure}
\centering
\includegraphics [width=0.5\textwidth]{system.eps}
\caption{A one-dimensional disordered chain with $N$ sites, represented in the figure by black dots, is coupled to a lead. Open circles represent sites of the lead. The outgoing plane wave is shown by the arrow where $0<\Re\{\tilde{k}\}<\pi$ and $\Im \{\tilde{k}\}<0$, so that it propagates left in the lead and its amplitude grows in the lead.}
\label{System}
\end{figure}
\section{Model and Its Effective Hamiltonian}
A discrete tight-binding one dimensional chain of length $N$ (shown by positive integers, $n=1,\,2,...,\,N$, used for indexing the sites of the chain in Fig. \ref{System}) is connected to an outer world (represented by a perfect lead which sites are shown by a zero and negative integers, $n=0,\,-1,\,-2,...$). Each site of the chain has the site energy $\epsilon_{n}$ where $\epsilon_{n}$ are statistically independent random variables chosen from some symmetric distribution. Each nearest neighbor site of the chain as well as of the lead is coupled by a hopping amplitude $t$. The hopping amplitude for the pair $n=0$ and $n=1$ is $t'$ which takes values from $t'=0$ (closed chain) to $t'=t$ (fully coupled chain). With this hopping, a particle, which is initially located somewhere in the chain, eventually escapes to the outer world.
Now we write down the Schr\"{o}dinger equation for the entire system,
\begin{eqnarray}
\label{Sch1}
-t\psi_{n+1}-t\psi_{n-1}&=&\tilde{\mathcal{E}}\psi_{n},~~~~~~~~\text{for $n<0$},
\\
\label{Sch2}
-t\psi_{-1}-t'\psi_{1}&=&\tilde{\mathcal{E}}\psi_{0}, ~~~~~~~~\text{for $n=0$},
\\
\label{Sch3}
-t'\psi_{0}-t\psi_{2}+\epsilon_{1}\psi_{1}&=&\tilde{\mathcal{E}}\psi_{1},~~~~~~~~\text{for $n=1$},
\\
\label{Sch4}
-t\psi_{n-1}-t\psi_{n+1}+\epsilon_{n}\psi_{n}&=&\tilde{\mathcal{E}}\psi_{n},~~~~~~~~\text{for $2\leq n\leq N$.}
\end{eqnarray}
In order to avoid cluttering of notations we always represent quantities corresponding to disordered system by {\it script letters} while quantities for the clean system are represented in usual math notations. Tilde is used to discriminate the open system case from the closed one. Equation (\ref{Sch1}) is for the lead where $\epsilon_{n}=0$. Equations (\ref{Sch2}, \ref{Sch3}) describe the lead-chain coupling and Eq. (\ref{Sch4}) is for the chain. As in, \cite{KS:08} we solve Eqs. (\ref{Sch1}-\ref{Sch4}) with a boundary condition of an outgoing plane wave in the lead, i.e., $\psi_{n_{\leq 0}}\propto \exp(-i\tilde{k}n)$ where $0<\Re\{\tilde{k}\}<\pi$ and $\Im \{\tilde{k}\}<0$. The condition on $\Re\{\tilde{k}\}$ ensures that the outgoing wave propagates to left in the lead. The condition on $\Im \{\tilde{k}\}$ is considered so that the amplitude of the resonance wave function grows in the lead. It comes from Eq. (\ref{Sch1}) that the complex energy $\tilde{\mathcal{E}}$ is related to the complex wave vector $\tilde{k}$ via the dispersion relation $\tilde{\mathcal{E}}=-2 t \cos(\tilde{k})$. Now we eliminate all $\psi_{n}$ for $n<1$ from Eqs. (\ref{Sch1}-\ref{Sch4}) and obtain
\begin{equation}\label{reduced}
-t\psi_{n+1}-t\psi_{n-1}+\tilde{\epsilon}_{n}\psi_{n}=\tilde{\mathcal{E}}\psi_{n},
\end{equation}
where
\begin{equation}\label{energy}
\tilde{\epsilon}_{n}=\epsilon_{n}-t\eta\, \exp(i\tilde{k})\delta_{n1},
\end{equation}
for $n=1,\,2,...,\,N$. The parameter $\eta=(t'/t)^2$ measures the coupling strength to the outside world.
An effective Hamiltonian defined by the Eq. (\ref{reduced}) is non-Hermitian. For instance, if $\mathcal{H}$ is the $N\times N$ tridiagonal Hermitian matrix which represents the Hamiltonian of the closed-disordered chain then one may write the effective Hamiltonian, $\tilde{\mathcal{H}}$, as
\begin{equation}\label{Hamil}
\tilde{\mathcal{H}}=\mathcal{H}-t\,\eta\,\lambda(\tilde{k})\, P.
\end{equation}
Here $P=|1\rangle\langle1|$ is the projection for site $n=1$ and $\lambda=\exp(i\tilde{k})$. The above non-Hermitian effective Hamiltonian has been first obtained by Terraneo and Guarnery \cite{GG:00}. The underlying result here is that the same relation (\ref{Hamil}) is valid for any Hermitian $\mathcal{H}$ representing a (closed) quantum system \cite{JF:10} which has $N$-dimensional state space. Resonances are characterized by the complex eigenvalues, $\tilde{\mathcal{E}}_{\alpha}$, of $\tilde{\mathcal{H}}$.
Note here dependency of $\tilde{\mathcal{H}}$ on the complex wave vector $\tilde{k}$ which is related to the complex energies via the dispersion relation mentioned above - this is not a standard eigenvalue problem. To standardize this problem ``parametric resonances'' are often used as an alternative. In this approach the dependence of $\lambda$ on $\tilde{k}$ is typically neglected, reducing thereby the problem of finding the eigenvalues of the effective Hamiltonian at chosen value of $\tilde{k}$. As expected, parametric resonances yield approximate statistical results which are close to those for the exact resonances in strongly localized regime \cite{GG:00}. Parametric resonances depend on a chosen parameter, for instance let $\tilde{k}=k_{0}$ and we fix it in the middle of the energy band, $k_{0}=\pi/2$. Writing explicitly
\begin{equation}
\lambda(\tilde{k})=
\begin{cases}
\exp(i \tilde{k}), & \text {for exact resonances,}
\\
i, & \text {for parametric resonances.}
\end{cases}
\end{equation}
From now on we set the energy scale by taking $t=1$, denoting the complex variable $\tilde{\mathcal{E}}/t$ by $\tilde{\mathcal{Z}}$. We denote the Hamiltonian matrix representing the closed-clean chain by $H$. It differs from $\mathcal{H}$ only at the diagonal as, for the clean chain case, all the site energies are zero. Calculation of the eigenvalues of $H$ is a standard exercise where one derives $z_{\alpha}=-2 \cos[\alpha\pi/(N+1)]$ for $\alpha=1,...,N$.
Before going into a detail treatment to the problem, we should first sketch the outline of our approach. We are interested in a weak disorder regime. Since our approach rely on perturbation theory, we need complex energies of open-clean chain, i.e., the $\tilde{z}_{\alpha}$s. So we will begin with calculating the resonances for open-clean chain of finite length. Then we will do the perturbation series expansion up to the second order of strength of the disorder. This will be followed by the derivation of the statistical results. Finally, we will consider the large-$N$ limit of these results.
\section{Open-clean Chain}
We begin with defining the resolvent $\tilde{\mathcal{G}}(z)=(z-\tilde{\mathcal{H}})^{-1}$. Using Eq. (\ref{Hamil}) we may also write
\begin{equation}\label{resolvent}
\tilde{\mathcal{G}}(z)=(z-\mathcal{H}+\eta\,\lambda\,P)^{-1}.
\end{equation}
For the open-clean chain we define the resolvent
\begin{eqnarray}
\tilde{G}(z)&=&(z-H+\eta\,\lambda\,P)^{-1}
\nonumber
\\
&=&
(1+\eta\,\lambda\,GP)^{-1}\,G,
\end{eqnarray}
where we have introduced $G(z)=(z-H)^{-1}$ as the resolvent for the ``unperturbed" closed-clean chain. Resonances correspond to the singularities of the matrix $\tilde{G}_{mn}(z)$, or to the roots of the secular equation
\begin{equation}\label{charcpol}
F(z)=0
=1+\eta\lambda G_{11}(z),
\end{equation}
where $G_{11}$ is the $\{1,\,1\}$ element of the matrix $G$ in site representation. ($G_{nm}(z)=\langle n|(z-H)^{-1}|m\rangle$.)
\begin{figure}[!t]
\centering
\includegraphics [width=0.75\textwidth]{ansatz-exact.eps}
\caption{Comparison of the result (\ref{resz}) (pluses) with the numerical solution of the polynomial equation (\ref{fincharpol}) (circles, squares and diamonds) for the exact resonances where $\eta=0.5,\,0.81$ and $0.99$. We have considered $N=100$.}
\label{ansatz-exact}
\end{figure}
To obtain $G_{11}$ for finite $N$, we use the ordinary difference equation (ODE),
\begin{equation}\label{ODE}
\psi_{n+1}+\psi_{n-1}+z\,\psi_{n}=0,
\end{equation}
with the boundary conditions $\psi_{0}=\psi_{N+1}=0$. This equation is obtained from Eq. (\ref{reduced}) by setting all $\epsilon_{n}=0$. Next we consider $u_{n}(z)$ and $v_{n}(z)$ to be the two linearly independent functions which satisfy the ODE
\begin{eqnarray}\label{un}
u_{n+1}+u_{n-1}+z\,u_{n}&=&0,
\\
\label{vn}
v_{n+1}+v_{n-1}+z\,v_{n}&=&0,
\end{eqnarray}
where $u_{0}=v_{N+1}=0$. Since norm of $u_{n},\,v_{n}$ is arbitrary, we fix $u_{1}=v_{N}=1$. Further we claim that the resolvent is given by
\begin{equation}\label{G0uv}
G_{nm}=-\dfrac{u_{n}\,v_{m}\Theta(m-n)+u_{m}\,v_{n}\Theta(n-m)}{W_{n}}.
\end{equation}
Here $\Theta(n)$ is the unit-step function and $W_{n}=u_{n}v_{n-1}-u_{n-1}v_{n}$ is the Wronskian. Using Eqs. (\ref{un}, \ref{vn}) it is straight forward to see that the Wronskian is independent of $n$. One can also check that
\begin{equation}
G_{n+1m}+G_{n-1m}+z\,G_{nm}=\delta_{nm}.
\end{equation}
We now set $u_{n}=v_{N+1-n}$ to match the initial value problem (\ref{un}, \ref{vn}) to the boundary value problem (\ref{ODE}). We find
\begin{equation}\label{Gun}
G_{11}=-\dfrac{u_{N}}{u_{N+1}}.
\end{equation}
The ODE (\ref{un}) is satisfied by the Chebyshev polynomial of the second kind, $U_{m}(-z/2)$, defined as
\begin{equation}
U_{m}(x)=\dfrac{\sin[(m+1)\cos^{-1}(x)]}{\sin[\cos^{-1}(x)]},
\end{equation}
for $U_{0}(x)=1$ and $U_{1}(x)=2x$. Since we have fixed $u_{1}=1$, therefore $u_{n}=U_{n-1}$, thus we can write Eq. (\ref{Gun}) as
\begin{equation}\label{GF}
G_{11}=-\dfrac{U_{N-1}(-z/2)}{U_{N}(-z/2)}=-\dfrac{\sin[N\,k]}{\sin[(N+1)k]}.
\end{equation}
Here the last equality follows from the energy dispersion relation. Using Eq. (\ref{GF}) in Eq. (\ref{charcpol}), we end up with an algebraic equation
\begin{equation}\label{fincharpol}
F(z)
=0=
1-\eta\lambda\dfrac{U_{N-1}(-z/2)}{U_{N}(-z/2)}.
\end{equation}
\begin{figure}
\centering
\includegraphics [width=0.75\textwidth]{ansatz-parametric.eps}
\caption{Repeated on the same pattern of Fig. \ref{ansatz-exact} but for parametric resonances.}
\label{ansatz-parametric}
\end{figure}
Zeros of $F(z)$ are the roots of a polynomial of order $N$. For exact resonances Eq. (\ref{fincharpol}) can be easily transformed into
\begin{equation}\label{Az}
[\mathsf{a}(z)]^{(2N+1)}=
\dfrac{\mathsf{a}(z)^{-1}-\eta\,\mathsf{a}(z)}{1-\eta},
\end{equation}
where
\begin{equation}\label{Aexp}
\mathsf{a}(z)=-\exp[ik(z)].
\end{equation}
In order to solve Eq. (\ref{Az}), we propose an ansatz assuming that opening of the system at one end causes $\mathcal{O}(N^{-1})$ complex corrections to the $k_{\alpha}$'s. Let
\begin{equation}\label{ansatz}
\tilde{k}_{\alpha}=k_{\alpha}+\dfrac{\Phi_{\alpha}}{N},
\end{equation}
where $\Phi_{\alpha}$ is a complex quantity and $k_{\alpha}=\alpha\pi/(N+1)$. Inserting this ansatz into Eqs. (\ref{Az}, \ref{Aexp}) we obtain
\begin{equation}\label{res}
\tilde{k}_{\alpha}= k_{\alpha}-
\dfrac{i}{2N}\,
\text{ln}\left[\Omega(k_{\alpha};\eta)\right]+\mathcal{O}\left(\dfrac{1}{N^2}\right),
\end{equation}
where
\begin{equation}\label{Omega}
\Omega(k;\eta)=
\dfrac{1-\eta \,e^{2i\,k_{\alpha}}}{1-\eta}
\end{equation}
Now, up to $\mathcal{O}(N^{-1})$, $\tilde{z}_{\alpha}$ may be written as
\begin{equation}\label{resz}
\tilde{z}_{\alpha}=-2\cos(k_{\alpha})-
\dfrac{i\sin(k_{\alpha})}{N}\,
\text{ln}(\Omega).
\end{equation}
The same result can be obtained for the parametric resonances, after repeating the similar steps, but with different $\Omega$:
\begin{equation}\label{Omgpara}
\Omega(k_{\alpha};\eta)=
\dfrac{1-i\eta \,e^{i\,k_{\alpha}}}
{1-i\eta \,e^{-i\,k_{\alpha}}}.
\end{equation}
\begin{figure}
\centering
\includegraphics [width=0.75\textwidth]{scatter-exact.eps}
\caption{Scatter plot for exact resonances where $N=100$, $\eta=0.81$. Dense points in the graph represent exact resonances in the disordered chain for $2500$ realizations where $W=0.015$. These are scattered around dots which represent exact resonances in the clean chain.}
\label{scatter_exact}
\end{figure}
One should bear in mind that there is no resonance for $\eta=1$, as the system is fully coupled to the lead. However, for parametric resonances, one artificially gets resonances even when $\eta=1$. Note that the result (\ref{resz}) is symmetric about the imaginary axis for both cases.
In Fig. \ref{ansatz-exact} and Fig. \ref{ansatz-parametric} we compare the numerical solutions of the polynomial equation (\ref{fincharpol}) with our results (\ref{resz}, \ref{Omega}, \ref{Omgpara}), for $N=100$, $\eta=0.50,\,0.81$ and $0.99$ and $N=100$, respectively for exact and parametric resonances. Eq. (\ref{fincharpol}) has been solved by using the Newton's method with the initial guess $\tilde{k}_{\alpha}=k_{\alpha}$. These figures show that our result (\ref{resz}) is close to the numerical solution. The agreement gets better as $\eta\rightarrow1$ (not shown here separately). However, the ansatz (\ref{ansatz}) is not valid near the band edges. Moreover, the agreement fails for parametric resonances near the middle of the band as $\eta\rightarrow 1$; see Fig. \ref{ansatz-parametric} for $\eta=0.99$.
\section{The Weak disorder limit}
\begin{figure}
\centering
\includegraphics [width=0.75\textwidth]{scatter-parametric.eps}
\caption{Scatter plot for parametric resonances where $N=100$, $\eta=0.81$ and $W=0.015$, for $5000$ realizations. As in Fig.\ref{scatter_exact}, here also dots represent the clean chain and points represent the disordered chain.}
\label{scatter_parametric}
\end{figure}
In the next stage of the problem we switch on a very weak disorder in the chain. From a second order perturbation theory we know that for a disordered infinitely long chain the localization length, $\xi(E)$, is maximum at the middle of the band. For small $W$ it is given by \cite{krammac,thou}
\begin{eqnarray}\label{xi}
\xi(E)=\dfrac{24(4t^2-E^2)}{W^2},
\end{eqnarray}
implying thereby, $\xi(0)=\dfrac{96 t^2}{W^2}$. However, the exact result shows a small deviation at the band center due to the breakdown of the second-order perturbation theory \cite{KWegner:1981}. We consider a limiting situation when $\xi(0)/N >> 1$. For instance, in Fig. \ref{scatter_exact} and in Fig. \ref{scatter_parametric}, we show the scatter plot ($\Re\{\tilde{\mathcal{Z}}_{\alpha}\}$ vs $\Im\{\tilde{\mathcal{Z}}_{\alpha}\}$) for exact and parametric resonances respectively. In both cases we have considered $N=100$, $\eta=0.81$ and $W=0.015$ so that $\xi(0)>>N$. As seen in these figures, complex energies of the disordered chain are scattered around the $\tilde{z}_{\alpha}$s.
We now calculate the corrections to $\tilde{z}_{\alpha}$ for such weak disorder case. It is suggestive here to deal with the self-energy. Let $\mathcal{S}_{1}(\epsilon_{2},...,\epsilon_{N};z)$ be the self-energy for the first site, defined via
\begin{eqnarray}\label{GS}
\mathcal{G}_{11}(z)=\dfrac{1}{z-\epsilon_{1}-\mathcal{S}_{1}(\{\epsilon\};z)}.
\end{eqnarray}
Here $\{\epsilon\}$ denotes the set $\epsilon_{2},...,\epsilon_{N}$ and $\mathcal{G}_{11}$ is the $\{1,\,1\}$ element of the resolvent $\mathcal{G}(z)=(z-\mathcal{H})^{-1}$, defined for the Hermitian matrix $\mathcal{H}$. For the later convenience we write
\begin{eqnarray}\label{Hborn}
\mathcal{H}=H+\mathcal{W},
\end{eqnarray}
where $\mathcal{W}=\sum_{\ell=1}^{N}\epsilon_{\ell}P_{\ell}$ and $P_{\ell}=|\,\ell\,\rangle\langle\,\ell\,|$ is the projection for the $\ell$'th site.
In the rest of the paper we will work out results only for the exact resonances. For the parametric resonance theses results can be carried out following similar steps, so we skip all the intermediate steps merely by stating the result at the end.
As before in Eq. (\ref{charcpol}), for disordered chain, resonances correspond to the roots of the secular equation
\begin{equation}\label{secular}
\mathcal{F}(z)=0=z-\epsilon_{1}-\mathcal{S}_{1}(\{\epsilon\};z)+\lambda\eta.
\end{equation}
Preserving $\tilde{z}_{\alpha}$ as the roots of Eq. (\ref{charcpol}), we define $\tilde{\mathcal{Z}}_{\alpha}$ as the roots of Eq. (\ref{secular}). Now we expand the roots $\tilde{\mathcal{Z}}_{\alpha}=\tilde{z}_{\alpha}+(\delta_{1} \tilde{\mathcal{Z}}_{\alpha})+(\delta_{2} \tilde{\mathcal{Z}}_{\alpha})$, assuming that $(\delta_{1}\tilde{\mathcal{Z}}_{\alpha})$ are linear while $(\delta_{2} \tilde{\mathcal{Z}}_{\alpha})$ are quadratic in the $\epsilon_{j}$ , for $j=1,...,N$. Then for $\mathcal{S}_{1}(\{\epsilon\};\tilde{\mathcal{Z}}_{\alpha})$, up to $\mathcal{O}(\{\epsilon\}^{2})$, we get
\begin{eqnarray}\label{Selfenergy}
\mathcal{S}_{1}(\{\epsilon\};\tilde{\mathcal{Z}}_{\alpha})
&=&
S_{1}(\{0\};\tilde{z}_{\alpha})+
\sum_{n=2}^{N}\epsilon_{n}
\left(\dfrac{\partial \mathcal{S}_{1}(\{\epsilon\};z)}{\partial\epsilon_{n}}\right)_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}
\nonumber
\\
&+&
(\delta_{1} \tilde{\mathcal{Z}}_{\alpha}+\delta_{2} \tilde{\mathcal{Z}}_{\alpha})\left(\dfrac{\partial \mathcal{S}_{1}(\{\epsilon\};z)}{\partial z}\right)_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}
\nonumber
\\
&+&
\dfrac{1}{2}\sum_{n,m=2}^{N}\epsilon_{n}\epsilon_{m}
\left(\dfrac{\partial^{2} \mathcal{S}_{1}(\{\epsilon\};z)}{\partial\epsilon_{n}\partial\epsilon_{m}}\right)_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}
\nonumber
\\
&+&
\dfrac{1}{2}(\delta_{1} \tilde{\mathcal{Z}}_{\alpha})^{2}\left(\dfrac{\partial^{2} \mathcal{S}_{1}(\{\epsilon\};z)}{\partial^{2} z}\right)_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}.
\end{eqnarray}
We will use this expansion in Eq. (\ref{secular}). Before that we evaluate
\begin{eqnarray}\label{SGdz}
1-\left(\dfrac{\partial \mathcal{S}_{1}(\{\epsilon\};z)}{\partial z}\right)_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}
=
\dfrac{\partial}{\partial z} \dfrac{1}{G_{11}(z)}\bigg|_{z=\tilde{z}_{\alpha}},
\end{eqnarray}
and,
\begin{eqnarray}\label{dGdz}
\dfrac{\partial \mathcal{S}_{1}(\{\epsilon\};z)}{\partial \epsilon_{n}}\bigg |_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}
&=&
\dfrac{1}{\left(G_{11}\right)^2}\,
\dfrac{\partial \mathcal{G}_{11}}{\partial \epsilon_{n}}\bigg|_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}},
\nonumber
\\
\end{eqnarray}
for $n\geq2$. These equalities come from Eq. (\ref{GS}). Finally, we calculate derivatives of $\mathcal{G}_{11}$, at $\{\epsilon\}=0$ and $z=\tilde{z}_{\alpha}$ with respect to $\{\epsilon\}$ by using Eq. (\ref{Hborn}) for the {\it Born-series} expansion of $\mathcal{G}(z)$. We find
\begin{eqnarray}\label{dGdeps}
\dfrac{\partial \mathcal{G}_{11}}{\partial \epsilon_{n}}\bigg|_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}
&=&
G_{1n}G_{n1}\bigg|_{z=\tilde{z}_{\alpha}}.
\end{eqnarray}
Grouping all these, for the first order corrections, we obtain
\begin{eqnarray}\label{FPT1}
&&(\delta_{1}\tilde{\mathcal{Z}}_{\alpha})
=
\dfrac{\epsilon_{1}+\sum_{n=2}^{N}\epsilon_{n}
\dfrac{G_{1n}G_{n1}}
{\left[G_{11}\right]^{2}}\Bigg|_{z=\tilde{z}_{\alpha}}}
{\dfrac{\partial}{\partial z} \dfrac{1}{G_{11}(z)}\bigg|_{z=\tilde{z}_{\alpha}}
+
\dfrac{i\eta\exp[i\tilde{k}_{\alpha}]}
{2\sin(\tilde{k}_{\alpha})}
}.
\end{eqnarray}
Similarly for the second order corrections we get
\begin{eqnarray}\label{dz2}
(\delta_{2} \tilde{\mathcal{Z}}_{\alpha})
&=&
\Bigg[\sum_{n,m=2}^{N}\epsilon_{n} \epsilon_{m}
\left\lbrace
\dfrac{G_{1n}G_{nm}G_{m1}}{[G_{11}]^{2}}
-
\dfrac{[G_{1n}G_{1m}]^{2}}
{[G_{11}]^{3}}
\right\rbrace
\nonumber
\\
&-&
\dfrac{(\delta_{1} \tilde{\mathcal{Z}}_{\alpha})^{2}}{2}
\left\lbrace\left(\dfrac{\partial^{2}}{\partial^{2} z}\dfrac{1}{G_{11}}\right)+\eta\left(\dfrac{d^{2}\exp(ik(z)}{d^{2}z}\right)\right\rbrace
\Bigg]_{\{\epsilon\}=0,z=\tilde{z}_{\alpha}}
\nonumber\\
&\times &
\Bigg[{\dfrac{\partial}{\partial z} \dfrac{1}{G_{11}(z)}\bigg|_{z=\tilde{z}_{\alpha}}
+
\dfrac{i\eta\exp[i\tilde{k}_{\alpha})}
{2\sin(\tilde{k}_{\alpha})}
}\Bigg]^{-1}.
\nonumber\\
\end{eqnarray}
Note that $(\delta_{1} \tilde{\mathcal{Z}}_{\alpha})$ and $(\delta_{2} \tilde{\mathcal{Z}}_{\alpha}) $ have been obtained in terms of the resolvent of the closed-clean chain which we already know in terms of Chebyshev polynomials; see Eq. (\ref{G0uv}) and the relation between $u_{n}$ and $v_{n}$ with Chebyshev polynomials.
\section{Statistics of the Scattered Complex Energies}
We are interested in the statistics of the scattered complex energies. For instance, using the first order result (\ref{FPT1}) of the perturbation theory, we calculate average of square of absolute shift in complex energies defined as, $\langle|(\Delta \tilde{\mathcal{Z}}_{\alpha})|^2\rangle\equiv\langle|(\tilde{\mathcal{Z}}_{\alpha}-\tilde{z}_{\alpha})|^2\rangle$. The angular brackets are used here to represent the averaging over many realizations of set of all random site energies $\{\epsilon_{n}\}$. This quantity gives a statistical account for the scattered complex energies. We also calculate $\langle\,( \Re\{\Delta\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$ and $\langle\,(\Im\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$, viz, average of square of the real and the imaginary part of the shift $(\tilde{\mathcal{Z}}_{\alpha}-\tilde{z}_{\alpha})$, respectively. To obtain the latter quantities we need first to calculate $\langle\,(\Delta \tilde{\mathcal{Z}}_{\alpha})\,^{2}\rangle$ and $\langle\,[(\Delta \tilde{\mathcal{Z}}_{\alpha})^{*}]^{2}\,\rangle$, since
\begin{eqnarray}
(\Re\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}=\dfrac{(\Delta \tilde{\mathcal{Z}}_{\alpha})\,^{2}+[(\Delta \tilde{\mathcal{Z}}_{\alpha})^{*}]^{2}+2(|\Delta \tilde{\mathcal{Z}}_{\alpha})\,|^{2})}{4},
\nonumber
\\
\\
(\Im\{\Delta\tilde{\mathcal{Z}}_{\alpha}\})^{2}=-\dfrac{(\Delta \tilde{\mathcal{Z}}_{\alpha})\,^{2}+[(\Delta \tilde{\mathcal{Z}}_{\alpha})^{*}]^{2}-2(|\Delta \tilde{\mathcal{Z}}_{\alpha})\,|^{2})}{4}.
\nonumber
\\
\end{eqnarray}
Here we have used $\{^{*}\}$ to represent the complex conjugate (c.c.).
For all these three statistics we simplify $(\delta_{1}\tilde{\mathcal{Z}}_{\alpha})$, given in Eq. (\ref{FPT1}), in terms of Chebyshev polynomials as
\begin{eqnarray}\label{FPT2}
(\delta_{1}\tilde{\mathcal{Z}}_{\alpha})
&=&
\Bigg[\dfrac{\tilde{z}_{\alpha}^2-4}{2}
\sum_{n=1}^{N}\epsilon_{n}
\left(U_{N-n}(\tilde{z}_{\alpha}/2)\right)^{2}
\Bigg]
\nonumber\\
&\times&
\Bigg[U_{N-1}(\tilde{z}_{\alpha}/2)T_{N+1}(\tilde{z}_{\alpha}/2)-N
\nonumber\\
&-&
i\eta\exp[i\tilde{k}_{\alpha}]\sin(\tilde{k}_{\alpha})[U_{N-1}(\tilde{z}_{\alpha}/2)]^2
\Bigg]^{-1},
\end{eqnarray}
where $T_{m}(z)=\cos[m\cos^{-1}(z)]$ is the Chebyshev polynomial of the first kind. Further simplifications occur when these polynomials are expressed in their trigonometric forms. For instance, let's calculate $|(\Delta \tilde{\mathcal{Z}}_{\alpha})|^2$, with $ \tilde{z}_{\alpha}/2=\cos( \tilde{\theta}_{\alpha})$ where $\tilde{\theta}_{\alpha}=\pi- \tilde{k}_{\alpha}$. We obtain
\begin{eqnarray}\label{abdz2}
\dfrac{|(\Delta \tilde{\mathcal{Z}}_{\alpha})|^2}{4}
=
\dfrac{\sum_{n,m=1}^{N}\epsilon_{n'}\epsilon_{m'}\sin^{2}(n' \tilde{\theta}_{\alpha})\sin^{2}(m'\tilde{\theta}^{*}_{\alpha})}
{|D(\tilde{z_{\alpha}})|^{2}}.
\end{eqnarray}
Here $n'$ and $m'$ are respectively $N+1-n$ and $N+1-m$, and $D(\tilde{z_{\alpha}})$ is simply the quantity in the second bracket of Eq. (\ref{FPT2}). Averaging releases one of the summation as the $\epsilon_{j}'$s are statistically independent-identically-distributed (i.i.d.) random variables. We simply have
\begin{eqnarray}\label{abdz3}
\langle|(\Delta \tilde{\mathcal{Z}}_{\alpha})|^2\rangle
=
\sigma^{2}\dfrac{\sum_{n=1}^{N}4\sin^{2}(n\tilde{\theta}_{\alpha})\sin^{2}(n\tilde{\theta}^{*}_{\alpha})}
{|D|^{2}},
\end{eqnarray}
where $\sigma^{2}$ is variance of the $\epsilon_{j}'$s.
Summation in the above equality can be performed by using trigonometric identities. For instance, we first write
\begin{eqnarray}\label{abdz4}
4\sin^{2}(n\theta)\sin^{2}(n\theta^{*})
&=&
1-\cos(2n\theta)-\cos(2n\theta^{*})
\nonumber\\
&+&
\dfrac{\cos(4\,n\,\Re\{\theta\})+\cos(4\,i\,n\,\Im\{\theta\})}{2},
\end{eqnarray}
and we use the summation formula
\begin{eqnarray}\label{abdz5}
\sum_{n=1}^{N}\cos(n\theta)
&=&
\dfrac{1}{2}
\left[
\dfrac{\sin[(N+1/2)\theta]}{\sin(\theta/2)}-1
\right].
\end{eqnarray}
It turns out after some trigonometry that one can write the summation in a closed form. We find
\begin{eqnarray}
\sum_{n=1}^{N}4\sin^{2}(n\theta)\sin^{2}(n\theta^{*})
&=&N+\dfrac{1}{2}-\dfrac{U_{2N}+U_{2N}^{*}}{2}+
\dfrac{T^{*}_{2N+2}T_{2N}-T_{2N+2}T^{*}_{2N}}{2[T^{*}_{2}-T_{2}]}.
\nonumber
\\
\end{eqnarray}
Here the argument of the polynomials is $\tilde{z}_{\alpha}/2$ and for their complex conjugate it is $\tilde{z}^{*}_{\alpha}/2$. Finally, we write down finite-$N$ result for average of the absolute square of the shift,
\begin{eqnarray}
\label{abdz6}
\langle|(\Delta \tilde{\mathcal{Z}}_{\alpha})|^2\rangle
=
\sigma^{2}\dfrac{N+\dfrac{1}{2}-\dfrac{U_{2N}+U_{2N}^{*}}{2}
+
\dfrac{T^{*}_{2N+2}T_{2N}-T_{2N+2}T^{*}_{2N}}{2[T^{*}_{2}-T_{2}]}}
{|D|^{2}}.
\end{eqnarray}
We now turn our attention to large-$N$ behavior of the result (\ref{abdz6}). For this purpose we use the ansatz (\ref{ansatz}) and result the (\ref{res}) for $\tilde{k}_{\alpha}$. Large-N behavior for the Chebyshev polynomials, with argument $\tilde{z}_{\alpha}$, may be calculated as
\begin{eqnarray}
T_{2N}(\tilde{z}_{\alpha}/2)&=&\dfrac{\exp[2iN\tilde{\theta}_{\alpha}]+\exp[-2iN\tilde{\theta}_{\alpha}]}{2}
\nonumber
\\
&\approx&
\dfrac{\Omega({k}_{\alpha};\eta)\exp(-2i{k}_{\alpha})+\left[\Omega({k}_{\alpha};\eta)\right]^{-1}\exp(2i{k}_{\alpha})}{2}
\end{eqnarray}
\begin{eqnarray}
T_{2N+2}(\tilde{z}_{\alpha}/2)
=
\dfrac{\Omega(k_{\alpha};\eta)+\left[\Omega(k_{\alpha};\eta)\right]^{-1}}{2}+\mathcal{O}(N^{-1}),
\end{eqnarray}
\begin{eqnarray}
U_{2N}(\tilde{z}_{\alpha}/2)&=&\dfrac{\exp[i(2N+1)\tilde{\theta}_{\alpha}]-\exp[-i(2N+1)\tilde{\theta}_{\alpha}]}
{\exp(i\tilde{\theta}_{\alpha})-\exp(-i\tilde{\theta}_{\alpha})}
\nonumber
\\
&\approx&
\dfrac{\Omega({k}_{\alpha};\eta)\exp(-i{k}_{\alpha})-\left[\Omega({k}_{\alpha};\eta)\right]^{-1}\exp(i{k}_{\alpha})}{\exp(-i{k}_{\alpha})-\exp(i{k}_{\alpha})}.
\end{eqnarray}
Finally,
\begin{eqnarray}
T_{2}(\tilde{z}^{*}_{\alpha}/2)- T_{2}(\tilde{z}_{\alpha}/2)
&=& -\dfrac{2i\,\Im\{\Phi_{\alpha}\}}{N}\,
{z}_{\alpha}+\mathcal{O}(N^{-2})
\nonumber
\\
&\approx&
\dfrac{4i}{N}\, \cos(k_{\alpha})\,\Im\{\Phi_{\alpha}\},
\end{eqnarray}
where we have used the ansatz (\ref{ansatz}) in the second order polynomial $T_{2}(z)=2z^{2}-1$ and $\Im\{\Phi_{\alpha}\}=-\sin({k}_{\alpha})\,\text{ln}(|\Omega|)$, as obtained from Eqs. (\ref{ansatz}, \ref{res}).
We can now plug in these results in Eq. (\ref{abdz6}). These asymptotic results gives the numerator as ($N+a1+a2/(a3/N)$) where $a1,\,a2/a3$ are $\mathcal{O}(N^{0})$. Similarly we obtain denominator as ($\,N^{2}+b1\,N+b2$) where $b1$ and $b2$ are $\mathcal{O}(N^{0})$; see Appendix A for details. Thus in the leading order, we obtain
\begin{eqnarray}\label{abdz7}
\langle|(\Delta \tilde{\mathcal{Z}}_{\alpha})|^2\rangle
&=&
\dfrac{\sigma^{2}}{N}\,
\left(
1+\dfrac{1}{8}\,
\dfrac{
\left(
|\Omega|^{2}-|\Omega|^{-2}
\right)}
{\text{ln}(|\Omega|)}
\right).
\end{eqnarray}
\begin{figure}
\centering
\includegraphics [width=0.75\textwidth]{ex-finitevsLarge.eps}
\caption{Asymptotic results for $\langle\,|\delta_{1}\tilde{\mathcal{Z}}_{\alpha}|^{2}\,\rangle/\sigma^{2}$, $\langle\,(\Re\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$ and $\langle\,(\Im\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$, shown by filled circles, squares and diamonds, vs energy index $\alpha$. We have compared here the finite-$N$ results, shown by open circles, for the exact resonances where $N=100$ and $\eta=0.81$. In the set we show these results for 14 energy indices near the middle of the energy band but on a different scale.}
\label{ex_finitevsLarge}
\end{figure}
What follows next is the calculation of large-$N$ results for $\langle\,( \Re\{\Delta\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$ and $\langle\,(\Im\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$. Since we need first to calculate $\langle\,(\Delta \tilde{\mathcal{Z}}_{\alpha})\,^{2}\rangle$ and $\langle\,[(\Delta \tilde{\mathcal{Z}}_{\alpha})^{*}]^{2}\,\rangle$, from Eq. (\ref{FPT2}) we obtain after averaging
\begin{eqnarray}\label{dzsq}
\langle\,(\Delta \tilde{\mathcal{Z}}_{\alpha})^{2}\,\rangle
=\sigma^{2}\dfrac{\sum_{n=1}^{N}4\sin^{4}(n\tilde{\theta}_{\alpha})}
{D^{2}},
\end{eqnarray}
and
\begin{eqnarray}\label{cdzsq}
\langle\,[(\Delta \tilde{\mathcal{Z}}_{\alpha})^{*}]^{2}\,\rangle
=\sigma^{2}\dfrac{\sum_{n=1}^{N}4\sin^{4}(n\tilde{\theta}^{*}_{\alpha})}
{(D^{*})^{2}}.
\end{eqnarray}
\begin{figure}
\centering
\includegraphics [width=0.75\textwidth]{pa-finitevsLarge.eps}
\caption{Shown on the same pattern of Fig. \ref{ex_finitevsLarge} but for the parametric resonances where $N=500$ and $\eta=0.81$. In this figure $\langle\,|\delta_{1}\tilde{\mathcal{Z}}_{\alpha}|^{2}\,\rangle/\sigma^{2}$, $\langle\,(\Re\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$ and $\langle\,(\Im\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$ are shown respectively by pluses, crosses and stars. The inset is shown for the indices near the middle of the band.}
\label{pa_finitevsLarge}
\end{figure}
For summation we use the formula \cite{GR}
\begin{eqnarray}\label{sumsin4}
\sum_{n=1}^{N} \sin^{4}(n\,\theta)
&=&
\dfrac{1}{8}
\Bigg[
3N-\dfrac{\sin(N\theta)}{\sin(\theta)}
\big(4\cos[(N+1)\theta]
\nonumber
\\
&-&
\dfrac{\cos[2(N+1)\theta]\,\cos(N\theta)}{\cos(\theta)}
\big)\,
\Bigg],
\nonumber
\\
\sum_{n=1}^{N} \sin^{4}(n\,\tilde{\theta}_{\alpha})&=&
\dfrac{1}{8}
\left[
3N-4U_{N-1}T_{N+1}+\dfrac{T_{2N+2}U_{2N-1}}{\tilde{z}_{\alpha}}
\right],
\end{eqnarray}
where in the second equality the polynomials have argument $\tilde{z}_{\alpha}/2$ with $2\cos(\tilde{\theta}_{\alpha})=\tilde{z}_{\alpha}$. Similarly for the summation in Eq. (\ref{cdzsq}) one gets the polynomials with argument $\tilde{z}^{*}_{\alpha}/2$. One can now use the equality (\ref{sumsin4}) in Eqs. (\ref{cdzsq}, \ref{cdzsq}) in order to derive finite-$N$ result for $\langle\,(\Re\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$ and $\langle\,(\Im\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$. For large-$N$ we make use of the ansatz (\ref{ansatz}) and calculate the leading order contribution as
\begin{eqnarray}\label{LargeNdx}
\langle\,(\Re\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle
&=&
\dfrac{\sigma^{2}}{2N}
\Bigg\{\dfrac{5}{2}+\dfrac{1}{8}\,
\dfrac{
\left(
|\Omega|^{2}-|\Omega|^{-2}
\right)}
{\text{ln}(|\Omega|)}
+
g(k_{\alpha})\Bigg\},
\\
\nonumber
\\
\label{LargeNdy}
\langle\,(\Im\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle
&=&
-\dfrac{\sigma^{2}}{2N}
\Bigg\{\dfrac{1}{2}-\dfrac{1}{8}\,
\dfrac{
\left(
|\Omega|^{2}-|\Omega|^{-2}
\right)}
{\text{ln}(|\Omega|)}
+
g(k_{\alpha})\Bigg\},
\end{eqnarray}
where
\begin{eqnarray}
g(k_{\alpha})&=&
\dfrac{1}{8}
\left(
\dfrac{e^{-2ik_{\alpha}}\Omega^{2}-e^{2ik_{\alpha}}\Omega^{-2}}
{4i \sin(k_{\alpha})\left(2N\cos(k_{\alpha})+i\sin(k_{\alpha})\text{ln}(\Omega)\right)}
\right)
+
(\text{c.c.}).
\end{eqnarray}
Equations (\ref{abdz7}, \ref{LargeNdx}, \ref{LargeNdy}) are our main analytical results and they are also valid for parametric resonances with the $\Omega$ given in Eq. (\ref{Omgpara}). In Fig. \ref{ex_finitevsLarge} we verify the asymptotic results (\ref{abdz7}, \ref{LargeNdx}, \ref{LargeNdy}) against their finite-$N$ counterparts, for exact resonances with $N=100$ and $\eta=0.81$. Fig. \ref{pa_finitevsLarge} is repeated on the same pattern but for parametric resonances where $N=500$ and $\eta=0.81$. They confirm that the asymptotic results give a good account for the finite-$N$ results. However, there are some exception near the edges (not visible on the scale of the plot) where the ansatz (\ref{ansatz}) is not valid.
It turns out that in order to calculate the DOR we need the second order corrections $(\delta_{2}\tilde{\mathcal{Z}}_{\alpha})$, derived in Eq. (\ref{dz2}). We have followed the method used earlier \cite{FZ:99} for Hatano-Nelson Model \cite{HN:97}. However, we have not been able to obtain a closed expression of the DOR. This is discussed in Appendix B where we leave the calculations with a formal expression for the DOR.
\section{Numerical Methods and Verification of The Eqs. (\ref{abdz7}, \ref{LargeNdx}, \ref{LargeNdy})}
\begin{figure}
\centering
\includegraphics [width=0.75\textwidth]{ex-numericsvsLarge.eps}
\caption{Comparison of asymptotic results with numerics, for exact resonances where $N=100$, $W=0.015$ and $\eta=0.81$. In this figure, filled circles, squares and diamonds are the numerical results respectively for $\langle\,|\delta_{1}\tilde{\mathcal{Z}}_{\alpha}|^{2}\,\rangle/\sigma^{2}$, $\langle\,(\Re\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$ and $\langle\,(\Im\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$ while open circles are the rescaled theories (\ref{abdz7}, \ref{LargeNdx}) and (\ref{LargeNdy}). In the inset we show a comparison for 14 indices near the middle of the energy band on a different scale of the plot.}
\label{ex_numericsvsLarge}
\end{figure}
Numerical simulations for parametric resonance are always cost efficient. The reason being that there one deals with standard eigenvalue problem for which many fast subroutine packages are available, for instance LAPACK. On the other hand to verify the results (\ref{abdz7}, \ref{LargeNdx}, \ref{LargeNdy}) for exact resonances, where one needs to obtain numerical solutions of a characteristic polynomial equation of order $N$ in a complex plane, there is no as good algorithm. In this paper we show results for the exact resonances by calculating roots of the characteristic polynomial where we have used a cost efficient numerical subroutine {\it ezero}. The subroutine is available on the CPC program library. There is one major advantage of using this subroutine over other methods, for instance the Newton's method. This subroutine does not require initial guesses for the roots but only the contour which encloses all the roots of the polynomial. Besides, it also avoids calculating the derivatives which may result into numerical overflow.
In alternative to {\it ezero} we have used a different approach for calculating the roots. We survey the complex $\tilde{k}$-plane for the zeros of the $\text{Det}[M(\tilde{k})M(\tilde{k})^{\dagger}]$ where $M_{rs}=-2\cos(\tilde{k})\delta_{rs}-\tilde{\mathcal{H}}_{rs}$ for $r,s=1,...,N$ \cite{Neuberger}. (In our system $-\pi<\Re\{\tilde{k}\}<\pi$ and $\Im\{\tilde{k}\}<0$.) These zeros give the eigenvalues of $\tilde{\mathcal{H}}$. However, in the latter approach it is advisable to disintegrate the complex plane into small cells at first and then at every iteration into smaller one - only for $N$ cells which contain minima of the lowest eigenvalue and throwing the rest out. In this way one makes the algorithm faster and obtain the zeros in a reasonable precision. For a tridiagonal matrix this algorithm consumes a time which roughly grows with $N^3$. However, while comparing the two methods on a simple machine we find that the method used in {\it ezero} is much faster than the method described here. We refer to \cite{ezero} for further details of this subroutine.
In Fig. \ref{ex_numericsvsLarge}, we compare asymptotic results with simulation done for the total number of realizations $L=2500$, for exact resonances. In Fig. \ref{pa_numericsvsLarge} we compare numerical results obtained for parametric resonances, where $N=500$, $\eta=0.81$ and $L=5000$, with our theory for large-$N$. Though we have considered only the flat disorder yet our results are valid for the Gaussian or other symmetric distribution functions. These figures show that our asymptotic results are in fair agreement with the numerical results for almost all $\alpha$. For instance, near the middle of the energy band it describes reasonably well a dip and a peak, respectively in the $\langle\,(\Im\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$ and $\langle\,(\Re\{\Delta \tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle$. These two opposite effects, however, cancel out in $\langle\,(|\{\Delta \tilde{\mathcal{Z}}_{\alpha}\}|)^{2}\,\rangle$.
\begin{figure}
\centering
\includegraphics [width=0.75\textwidth]{pa-numericsvsLarge.eps}
\caption{ Shown on the same pattern of Fig. \ref{ex_numericsvsLarge} but for the parametric resonances where $N=500$, $W=0.015$ and $\eta=0.81$. These numerical results are obtained from the diagonalization of $N$-dimensional matrices for $5000$ realizations. In this figure $\langle\,|\delta_{1}\tilde{\mathcal{Z}}_{\alpha}|^{2}\,\rangle/\sigma^{2}$, $\langle\,(\Re\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$ and $\langle\,(\Im\{\delta_{1}\tilde{\mathcal{Z}}_{\alpha}\})^{2}\,\rangle/\sigma^{2}$ are shown respectively by pluses, crosses and stars while open circles are the rescaled theories (\ref{abdz7}, \ref{LargeNdx}) and (\ref{LargeNdy}).}
\label{pa_numericsvsLarge}
\end{figure}
\section{Conclusion}
In conclusion, we have studied resonances in a one dimensional discrete tight-binding open chain in a weak disorder limit. In this study we have calculated complex energies in an open-clean chain of finite length. The result we obtain is a polynomial equation which we have been able to solve for long chains using an ansatz for the solution. To the best of our knowledge, this result has never been derived before. We have used a perturbation theory up to the second order where we have derived the first and the second order corrections to the complex energies in terms of Chebyshev polynomials. The first order corrections have been useful to obtain closed form of the statistical results for the scattered complex energies. These results have been further simplified for long chains where we obtain compact results. The asymptotic results have been verified against numerics. Evidently, in the weak disorder limit the perturbation theory predicts nice statistical results. Our results are new in these studies and they could be useful in the further studies of such systems.
It would be interesting to study statistics of resonances in the weak disorder limit for higher dimensional models as well as for the cases when the site energies are not independent random variables but they are correlated with each other \cite{Izrailev:99}. Besides, there has been growing interest for the case when $M$ sites are connected to the outer world where $1\le M\le N$ \cite{Borgonovi:2012}. We believe that our methods could be useful for the study of such models. Finally, we mention the case where $\xi(0)\sim\mathcal{O}(N)$. It requires a separate investigation as our perturbative analysis fails in this limit.
\ack
The author is thankful to Boris Shapiro for suggesting the problem to him. The author would also like to give credit to Joshua Feinberg for the derivation of some of the equations in Secs. III and IV and also for the help in Appendix B. Discussions with both of them are gratefully acknowledged. The author also acknowledges Marko \v Znidari\v c and Thomas H. Seligman for reading the manuscript.
Support from ISF-1067 and generous hospitality of Technion Institute are also acknowledged. Additional support by the project 79613 by CONACyT, Mexico, is acknowledged.
|
\section{Introduction}
Crystalline quality, structural parameters and surface morphology of thin films are generally influenced by their thickness due to several factors such as strain, defect formation and change in growth mode. Such changes affect the physical properties of thin films. Hence, whenever new functional thin films are prepared, it always becomes immediately interesting to explore the effect of layer thickness on the structural and physical properties.
Among the family of newly discovered Fe-based superconducting compounds, increasing the layer thickness of FeSe and FeSe$_{0.5}$Te$_{0.5}$ thin films improves the superconducting transition temperature, $T_{\rm c}$, presumably due to lattice distortion by strain\,\cite{01,02}.
For Co-doped BaFe$_2$As$_2$ (Ba-122) films, only thickness studies regarding the buffer layers have been reported to date. Tarantini $et$ $al$ and Lee $et$ $al$ have found that single crystalline (La,Sr)(Al,Ta)O$_3$ substrates with 100 unit cells of epitaxial SrTiO$_3$ resulted in the highest $T_{\rm c}$ and the largest critical current density, $J_{\rm c}$\,\cite{03,04}. Even higher $T_{\rm c}$ and sharper out-of-plane and in-plane textures of the Fe/Ba-122 bilayers can be realized for 20\,nm thick epitaxial Fe buffer layers\,\cite{05}. However, no investigation of the effect of layer thickness of Co-doped Ba-122 on the structural and transport properties have been published to date. In this article, we report on the influence of layer thickness on the structural and transport properties of the Fe/Ba-122 bilayer system with a fixed Fe layer thickness.
\section{Experiment}
Epitaxial, smooth Fe buffer layers (20\,nm) were prepared by a two-step process, which involves a room temperature deposition of Fe on MgO (001) single crystalline substrates by pulsed laser deposition, PLD, followed by a high-temperature annealing at 750\,$^\circ$C, both in a UHV chamber (base pressure of 10$^{-10}$\,mbar). Prior to the Fe deposition, the substrate was heated to 1000\,$^\circ$C, held at this temperature for 30\,min, subsequently cooled to room temperature for cleaning. A KrF excimer laser (248\,nm) has been employed at a frequency of 5\,Hz for the deposition with an energy density of 3--5\,Jcm$^{-2}$ on the target. After the Fe buffer preparation, Co-doped Ba-122 layers were deposited at 750\,$^\circ$C with a laser repetition rate of 10\,Hz. Each deposition step was monitored by reflection high-energy electron diffraction, RHEED. The layer thickness, $d$, was varied in the range of 30\,nm to 225\,nm by controlling the number of laser pulses. Each layer thickness was confirmed by cross-sectional focused ion beam, FIB, cuts on multiple sample areas. The nominal composition of the PLD target was Ba:Fe:Co:As\,=\,1:1.84:0.16:2. The detailed target preparation can be found in reference\,\cite{06}. The phase purity of the target was determined by x-ray diffraction using a standard Bragg-Brentano geometry with Co-K$\alpha$ radiation. All the observed peaks were identified as Co-doped Ba-122. The lattice parameters refined via Rietveld analyses were $a=0.39586(2)$\,nm and $c=1.29825(6)$\,nm, respectively.
Surface morphology of the films was observed by atomic force microscopy, AFM. Out-of-plane texture and phase purity were investigated by x-ray diffraction in Bragg-Brentano geometry with Co-K$\alpha$ radiation. In-plane orientation of both Fe and Co-doped Ba-122 were investigated by using the 110 and 103 poles respectively in a texture goniometer operating with Cu-K$\alpha$ radiation. In order to evaluate the in-plane and out-of-plane lattice parameters of Co-doped Ba-122 precisely, high resolution reciprocal space maps, RSM, around the 109, 10\underline{11} and 11\underline{10} reflections were performed with Cu-K$\alpha$ radiation. Here, the 204 reflection of MgO was used as a reference to eliminate any errors by a misalignment of the substrate.
After the structural characterization, Au layers were deposited on the films by PLD at room temperature followed by ion beam etching to form bridges of 0.5\,mm width and 1\,mm length for transport measurements. Superconducting properties were measured in a Physical Property Measurement System (PPMS, Quantum Design) by a standard four-probe method with a criterion of 1\,$\rm\mu Vcm^{-1}$ for evaluating $J_{\rm c}$. In the angular-dependent $J_{\rm c}$ measurements, $J_{\rm c}(\Theta)$, the magnetic field, $H$, was applied in the maximum Lorentz force configuration ($H$ perpendicular to $J$) at an angle $\Theta$ measured from the $c$-axis. $T_{\rm c}$ is defined as 50\% of the normal state resistance at 30\,K.
\section{Results and discussion}
\begin{figure}[t]
\centering
\includegraphics[width=8.5cm]{Fig-1.pdf}
\caption{Representative RHEED images of Fe/Ba-122 bilayer ($d$=150\,nm). (a) MgO single crystalline substrate at room temperature after the heat treatment. Yellow arrows indicate the diffraction spots lie on the Laue circles. (b) Fe at room temperature at the end of deposition, (c) Fe at 750\,$^\circ$C, and (d) Co-doped Ba-122 at room temperature. The incident electron beam is along the MgO [110] azimuth.}
\label{fig:figure1}
\end{figure}
The diffraction pattern in the RHEED images of MgO substrate shows a series of spots lying on the Laue circles, indicative of a perfectly flat surface (figure\,\ref{fig:figure1}(a)). For the Fe buffer preparation, the RHEED images of Fe in figure\,\ref{fig:figure1}(b) confirm the epitaxial growth even at room temperature for $d$=150\,nm. The diffraction spots turn into streaks with increasing temperature (figure\,\ref{fig:figure1}(c)), indicative of smoothing of the surface\,\cite{07}. For Co-doped Ba-122, the diffraction patterns with long streaks centered at positions on the Laue circles is typical for a multilevel surface (figure\,\ref{fig:figure1}(d)); i.e. for a high number of smooth terraces separated by steps\,\cite{07}. Additionally, the spacing of the observed streaks indicates a surface reconstruction, which is consistent with the observation on single crystals reported in reference\,\cite{08}.
The AFM image of this film in figures\,\ref{fig:figure2}\,(a) and (b) further confirmed that the surface was flat with a root mean-square roughness, $R_{\rm rms}$, of 0.83\,nm. The AFM image also shows that Co-doped Ba-122 grows in the terraced-island mode with an average step height of 0.65\,nm, which is almost identical to half the lattice parameter $c$. All the films in this study have the same surface morphology, and their surface roughness are summarized in Table\,\ref{tab:table1}. However, the grains of the 30\,nm thick film are smaller and not well connected compared with the other films (figures.\,\ref{fig:figure2}(c) and (d)).
\begin{figure}
\centering
\includegraphics[width=8.5cm]{Fig-2.pdf}
\caption{(a) 2-Dimensional and (b) 3-dimensional AFM images (1$\mu$m$\times$1$\mu$m) of Fe/Ba-122 bilayer ($d$=150\,nm) exhibits a large number of terraced islands. (c), (d) The corresponding images of the 30\,nm thick film show that the grains are smaller and not well connected.}
\label{fig:figure2}
\end{figure}
The $\theta\rm/2\theta$-\,scans for the Fe/Ba-122 bilayers with different layer thickness do not show any secondary phases (figure\,\ref{fig:figure3}). The pronounced 00$l$ reflections of Co-doped Ba-122 together with the 002 reflection of MgO and Fe are observed for all films, indicating a $c$-axis orientation. For thickness $d\geq90\,\rm nm$, a 110 component is observed whose intensity becomes gradually stronger with increasing $d$. In addition, the ratio of the diffraction intensity for the 110 and 004 is increased with $d$ (Table\,\ref{tab:table1}). This 110 component is also visible in the 103 pole figure measurement, and its epitaxial relation to the substrate is (110)[001]Ba-122$\|$(001)[110]MgO and (110)[001]Ba-122$\|$(001)[$\overline{1}$10]MgO. It should be noted, however, that the amount of the 110 component is small since the ratio of the diffraction intensity for the 110 and 004 is less than 0.01 for all films. Here, the corresponding value for a randomly oriented grain is 3.28 (ICDD card number 01-077-6875). Indeed, this small amount of 110 component does not compromise the crystalline quality as shown in Table\,\ref{tab:table1}. The full width at half maximum (FWHM), $\Delta\omega$, of the 004 rocking curve and the average $\Delta\phi$ of the 103 reflection of Co-doped Ba-122 are getting smaller with increasing $d$.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{Fig-3.pdf}
\caption{The $\theta\rm/2\theta$-\,scans of Fe/Ba-122 bilayers with various layer thicknesses on (001) MgO substrates. Intensity of the 110 reflection is observed to increase with $d$.}
\label{fig:figure3}
\end{figure}
Another prominent feature is a shift of the 00$l$ reflections to lower angles with increasing $d$, indicating an increase in the lattice parameter $c$. Calculated lattice parameters $c$ of the films using the Nelson-Riley function are increasing from 1.274\,nm ($d=30\,{\rm nm}$) to 1.289\,nm ($d=225\,{\rm nm}$)\,\cite{09}, as shown in Table\,\ref{tab:table1}. 1. In this calculation, the 002 reflection is omitted in order to avoid excessive extrapolation. The correlation between the layer thickness and both the in-plane and the out-of-plane lattice parameters evaluated by RSM are exhibited in figures\,\ref{fig:figure4}\,(a) and (b). The lattice parameter $a$ is observed to decrease with $d$, while the out-of-plane lattice parameter behaves the opposite way. The evaluated lattice parameter $c$ by both methods (i.e. RSM and the $\theta\rm/2\theta$-\,scans) are almost identical for all the films within the experimental uncertainty, albeit high angle data of $2\theta\geq160\,^\circ$ are not measured in the $\theta\rm/2\theta$-\,scans. Since the FeAs tetrahedron in the Ba-122 bonds coherently to bcc Fe\,\cite{10}, the lattice parameter $a$ of a thin Co-doped Ba-122 layer is close to that of Fe multiplied by $\sqrt2$ (0.4053\,nm), which is slightly larger than $a$ of bulk Co-doped Ba-122 (i.e. PLD target), suggesting tensile strain in the film. The lattice parameter of Fe is almost constant at around 0.287\,nm confirmed by RSM. Here the respective lattice misfit of Fe/MgO and Co-doped Ba-122/Fe are -3.9\% and -2.4\%.
\begin{table}
\centering
\caption{\label{tab:table1}Surface roughness, average FWHM values of the $\phi$-scans and the $\omega$-scans, lattice parameter $c$ and diffraction intensity ratio, $I_{\rm 110}$/$I_{\rm 004}$, for Co-doped Ba-122 thin films with different layer thickness, $d$.}
\begin{tabular}{cccccc}
\br
$d$\,(nm)&$R_{\rm rms}$\,(nm)&$\Delta\omega(^\circ)$&$\Delta\phi(^\circ)$&$c$ (nm)&$I_{\rm 110}$/$I_{\rm 004}$\\
\hline
30&1.14&0.82&0.99&1.274&n.\,d.\\
70&0.85&0.96&1.26&1.285&n.\,d.\\
90&1.47& 0.65 & 0.88&1.284&$\leq 0.001$\\
150&0.83& 0.67 & 0.92&1.288&0.003\\
225&2.11& 0.60 & 0.81&1.289&0.007\\
\br
\end{tabular}
\end{table}
\begin{figure}
\centering
\includegraphics[width=8.5cm]{Fig-4.pdf}
\caption{(a) The lattice parameter $a$ is observed to decrease with $d$. (b) Correspondingly, the lattice parameter $c$ are increased with $d$. $a_{\rm Fe}$ is the lattice parameter of Fe. (c) The unit cell volume is almost constant with $d$ in the range of $30\,{\rm nm}\leq d\leq 90\,{\rm nm}$, while the thicker films deviate from this trend. Lines are guide to the eye.}
\label{fig:figure4}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=8.5cm]{Fig-5.pdf}
\caption{(a) The normalized resistive traces of the Fe/Ba-122 bilayers show a gradual increase in the $T_{\rm c}$ with increasing $d$. (b) The $T_{\rm c}$ of the Co-doped Ba-122 film is sensitive to the lattice distortion. Errors of the $T_{\rm c}$ is defined as a delta $T_{\rm c}$.}
\label{fig:figure5}
\end{figure}
The unit cell volume, $V=a^2c$, of Co-doped Ba-122 films is almost constant with $d$ in the range of $30\,{\rm nm}\leq d\leq 90\,{\rm nm}$, while the thicker films deviate from this trend presumably due to a relatively large contents of grain boundaries, GBs (figure\,\ref{fig:figure4}(c)). Nevertheless, the volume for all the films is larger than that of the bulk Co-doped Ba-122, which might be due to As deficiency. The correlation between the lattice parameters and As deficiency for Co-doped Ba-122 is not clear. However, both the in-plane and the out-of-plane lattice parameters of As-deficient LaFeAsOF are enlarged compared with that of the stoichiometric sample\,\cite{11}, resulting in a larger lattice volume of the As-deficient sample.
Shown in figure\,\ref{fig:figure5}\,(a) are the normalized resistive traces of the Co-doped Ba-122 with different $d$. The film with $d=70\,{\rm nm}$ shows the lowest $T_{\rm c}$ of around 18\,K. The 30\,nm thick film shows a broad transition width, which is a direct consequence of the small grain size together with poor connectivity. Another prominent feature is a clear jump of $T_{\rm c}$ between 70\,nm and 90\,nm thickness. From the $\theta\rm/2\theta$-\,scans in figure\,\ref{fig:figure3}, the 110 component is absent in the 70\,nm thick film whereas the 90\,nm thick film contained a small amount of the 110 grains. For $d \geq 90\,{\rm nm}$, the $T_{\rm c}$s are gradually improved up to 23\,K together with sharpening of the transition width by increasing $d$. It is clear from figure\,\ref{fig:figure5}\,(b) that $T_{\rm c}$ of the films improves with $c/a$, which is consistent with our previous results for films on different substrates\,\cite{06}.
\begin{figure}
\centering
\includegraphics[width=8.5cm]{Fig-6.pdf}
\caption{(a) The $E-J$ curves for the Co-doped Ba-122 film with $d=225\,\rm nm$ measured in various magnetic fields at 12.65\,K. Applied magnetic fields are parallel to the $c$-axis of the film. (b) The $J_{\rm c}(\Theta)$ of the films with various $d$ measured in an applied field of 1\,T at a reduced temperature of 0.538. The data of the 225\,nm thick film are not visible, since it shows the almost identical form of $J_{\rm c}(\Theta)$ as the 150\,nm thick film. (c) $J_{\rm c}(\Theta)$ presented fig.\,\ref{fig:figure6}\,(b) are normalized to the value at $\Theta=90^\circ$.}
\label{fig:figure6}
\end{figure}
The $E-J$ curves for the Co-doped Ba-122 film with $d=225\,\rm nm$ measured in various magnetic fields at 12.65\,K show a power-law relation, indicative of current limitation by depinning of flux lines rather than GB effects (figure\,\ref{fig:figure6}\,(a)). As stated earlier, thicker films contained a small amount of the 110 component. However, the film does not show any sign of weak-link behavior.
The $J_{\rm c}(\Theta)$ measured at a reduced temperature of $t=0.538$ ($t=T/T_{\rm c,0}$, where the $T_{\rm c,0}$ is onset temperature of zero resistance) are exhibited in figure\,\ref{fig:figure6}\,(b). All the films except the 30\,nm thick film can carry a high $J_{\rm c}$ of over $0.18\,{\rm MA/cm^2}$ in the whole angular range. The 30\,nm thick film shows one order of magnitude lower $J_{\rm c}(\Theta)$ than that of the other films, which is due to the small grain size together with poor connectivity. Indeed, $E-J$ curves of this film show a non-ohmic linear differential, NOLD, signature, indicative of $J_{\rm c}$ limitation by GBs\,\cite{12}.
Another prominent feature is a shift upward of $J_{\rm c}$ at $\Theta=180^\circ$ ($H$\,$\parallel$\,${c}$) as the thickness of Co-doped Ba-122 films increases. In particular a small $c$-axis angular peak is observed for $d\geq150\,\rm nm$ (fig.\,\ref{fig:figure6}\,(c)). These films show very similar forms of $J_{\rm c}(\Theta)$ with a low $J_{\rm c}$ anisotropy, $\gamma_{\rm J}=J_{\rm c}(90^\circ)/J_{\rm c}(180^\circ)$. For example, $\gamma_{\rm J}$ of the film with $d\geq150\,\rm nm$ is only around 1.2, whereas the corresponding values of the 90\,nm and 30\,nm thick films are 1.5 and 2, respectively. These results suggest that the GBs may contribute to the pinning along the $c$-axis\,\cite{13}, which opens the opportunity to tune $\gamma_{\rm J}$ by controlling the amount of the 110 textured component.
The implementation of very thin Fe buffer layers also yields the 110 textured component\,\cite{05}. However, the crystalline quality and the superconducting properties of Co-doped Ba-122 are compromised due to the presence of too many GBs. Hence, there may exist a threshold of the amount of GBs at which deterioration of the structural and superconducting properties sets in.
\section{Conclusion}
The effect of the superconducting layer thickness on the structural and superconducting properties of Co-doped Ba-122 films has been investigated. The texture quality and the superconducting transition temperature are improved by increasing the layer thickness due to stress relief. Increasing the layer thickness yields an additional 110 textured component, creating GBs. However, these GBs may constitute $c$-axis pinning centers within a certain amount, which leads to an increase in $J_{\rm c}$ at $H$\,$\parallel$\,${c}$ without compromising the structural and superconducting properties.
\ack
The authors thank J.\,Scheiter for help with FIB cut samples and E.\,Barbara for help with the AFM observation. We are also grateful to M.\,K\"{u}hnel and U.\,Besold for their technical support and S.\,F\"{a}hler for his RHEED software.
\section*{References}
|
\section{Introduction}
Ten years ago Ted Jacobson discovered Jacobson's Miracle Hair Growth Formula \cite{jacobson}, which shows how black holes can acquire scalar hair if they are embedded into regions in which the relevant scalar fields evolve slowly with time. We argue in this paper that this observation can significantly change the behaviour of black-hole binaries, and provide very robust new observational constraints on the existence of time-dependent scalar fields in the cosmological past. Unlike other recent efforts to constrain scalar-tensor theories using black hole physics \cite{axiverse}, our arguments do not rely on black-hole super-radiance.
Why consider light scalars? Light scalar fields provide one of the few known ways to consistently modify gravity over large distances, and one of the main ways in which candidate quantum theories of gravity can differ from one another is the spectrum of bosons they predict at very low energies. Such fields, if they exist and are sufficiently light, can mediate long-range forces that can observably compete with gravity. Because of this many of their properties have been carefully studied over the years \cite{damourrev, sctensrev, BD, otherlightsc, chameleon, galileon}.
Furthermore, scalars commonly arise in fundamental theories, although it is more unusual for them to be light enough to mediate forces over macroscopic distances. Scalars are rarely this light because quantum corrections famously tend to give large contributions to scalar masses, even if they would have been massless at the purely classical level. But in some circumstances symmetries can protect against masses, such as if the scalar is a pseudo-Goldstone boson \cite{pGB} for a spontaneously broken approximate symmetry. This is what protects the mass of an axion \cite{axionorig,axionrev}, or other axion-like fields \cite{moreaxion}. Alternatively, light scalars can arise as part of the low-energy limit of an extra-dimensional model, with mass protection coming from extra-dimensional symmetries like higher-dimensional general covariance \cite{SLEDCosmo, ubernat}. In particular, extremely light scalars with masses of order the present-day Hubble scale naturally arise in this way in extra-dimensional proposals to address the cosmological constant problem \cite{SLED, SLEDrev}.
Once a light scalar is present it is very unlikely to be time-independent, since a small potential can easily drive a light scalar to roll. So it is natural to examine how the time evolution of ambient scalar fields can affect the properties of black holes and other astronomical objects. It is in this context that Jacobson's observation of scalar hair growth by black holes is so interesting.
We find in this paper that the scalar hair Jacobson predicts for black holes can have potentially measurable implications, by modifying the post-Newtonian response of orbiting black holes. In particular, we find it provides an unexpected source of scalar dipole radiation for an orbiting black hole, which would have been unavailable if the black hole had no hair (as is usually supposed).
Specifically, for canonically normalized scalar fields that approach $\phi \to \phi_\infty + \mu t$ asymptotically far from a pair of orbiting near-Newtonian black holes (with masses $M_{\scriptscriptstyle A}$ and $M_{\scriptscriptstyle B}$), we find the fraction of power radiated into scalar fields to be proportional to $\Delta a^2$, where $\Delta a = 4 G \mu (M_{\scriptscriptstyle A} - M_{\scriptscriptstyle B})$. We apply this expression to the quasar OJ287, which has been interpreted as a double black hole binary, leading to a bound $|\, \mu| < (16 \, \hbox{days})^{-1}$ on the instantaneous time-variation of {\em any} sufficiently light (see below for precisely how light) scalar fields at redshift $z \simeq 0.3$.
Others \cite{OtherJMHGF} have also sought useful ways to apply Jacobson's Miracle Hair Growth Formula, through its potential implications for black-hole accretion \cite{BHaccretion}. Rather than using (as we do) the scalar hair to argue for dipole scalar radiation, these authors argue that in quintessence models primordial black holes can accumulate significant amounts of the quintessence field, potentially leading to observable effects in gravitational radiation.
This paper is organized as follows. In section \ref{sec:sctensbhb}, scalar-tensor gravity is introduced, and the no-hair theorem is explicitly illustrated for the static spherically-symmetric solutions of the vacuum field equations. In section \ref{sec:massschair}, the mass $M$ and scalar charge $Q$ of an astrophysical body are defined, in terms of both the behaviour of the fields asymptotically far from the body, and the functional derivatives of the point-particle action of the effective field theory in which the internal structure of the body has been integrated out. In section \ref{sec:bh}, Jacobson's time-dependent black hole solution with scalar hair (the Miracle Hair Growth Formula) is described. An analysis of the appropriate point-particle action is carried out, and it is found that this type of black hole couples to the scalar field differently than a star with the same scalar charge, implying that the standard formulas in the scalar-tensor literature for the Post-Keplerian orbital corrections need not apply to a binary system of two such black holes. Section \ref{sec:bhbinfr} then argues that standard formulae {\em do} apply for the leading-order rate of scalar dipole radiation emitted by a such a binary system. In section \ref{sec:oj287}, the result of this calculation is applied to the quasar OJ287. Finally, section \ref{sec:summary} summarizes our results and prospects for future work.
\section{Black holes: hair and dipole radiation}
\label{sec:bhhairdip}
This section contains the main formal results: describing black holes in scalar-tensor theories, with both static and slowly varying asymptotic scalar configurations. It also briefly summarizes the standard formulae for the dipole and quadrupole radiation rate from binary pairs in the near-Newtonian limit.
\subsection{Scalar-Tensor gravity and black-hole balding}
\label{sec:sctensbhb}
The simplest variants of scalar-tensor gravity involve a single scalar field, $\psi$, and the space-time metric, $\hat g_{\mu\nu}$, with action: $S = S_g + S_{\rm mat}$. Here $S_{\rm mat}$ describes how the fields $\psi$ and $g_{\mu\nu}$ couple to other matter kinds of matter and $S_g$ describes the mutual interactions of $\psi$ and the metric. Writing $S_g$ in a derivative expansion and keeping only up to two derivatives, the most general expression is
\begin{equation}
\label{actionv0.9}
S_{g} = - \int {\hbox{d}}^{4}x
\sqrt{-\hat g} \; \left\{ U(\psi) + \hat g^{\mu \nu}
\left[ \mathcal{F}(\psi) \, \hat{\mathcal{R}}_{\mu \nu}
+ \mathcal{G}(\psi) \, \partial_{\mu} \psi \,
\partial_{\nu} \psi \right] + \cdots \right\} \,,
\end{equation}
where the form taken by the functions $U(\psi)$, $\mathcal{F}(\psi)$ and $\mathcal{G}(\psi)$ characterize a particular scalar-tensor theory. Here $\hat{\mathcal{R}}_{\mu\nu}$ denotes the Ricci tensor\footnote{Conventions: we use metric signature $(-+++)$ and Weinberg's curvature conventions \cite{Wbg} (differing from MTW \cite{MTW} only by an overall sign for the Riemann tensor). Unless specified otherwise, units are chosen throughout with $\hbar = c = k_{\scriptscriptstyle B} = 1$.} constructed from the metric $\hat g_{\mu\nu}$. Of particular interest are scalar-tensor theories which satisfy the weak equivalence principle. This can be ensured by requiring that matter couplings, $S_{\rm mat}$, depend only on $\hat g_{\mu\nu}$ and not separately on $\psi$.
In practice the functions $\mathcal{F}$ and $\mathcal{G}$ need not be specified in detail because they can be dramatically simplified by performing appropriate field redefinitions of the form $\hat g_{\mu\nu} = A^{2}(\psi) \, g_{\mu\nu}$ and $\psi = \psi(\varphi)$, with the result
\begin{equation}
\label{action}
S_{g} = -\frac{1}{16 \pi G}\int {\hbox{d}}^{4}x \sqrt{-g}
\; \Bigl[ g^{\mu \nu} \left( \mathcal{R}_{\mu \nu}
+ 2 \, \partial_{\mu} \varphi \,
\partial_{\nu}\varphi \right) + 16 \pi G \, V(\varphi)
+ \cdots \Bigr] \,,
\end{equation}
where $G$ is Newton's constant. This leaves $V(\varphi)$ as the only important function governing the long-distance properties of the scalar-gravity couplings. For those theories for which $S_{\rm mat}$ depends only on $\hat g_{\mu\nu} = A^2(\varphi) g_{\mu\nu}$ and not separately on $\varphi$, all of the direct matter couplings of the field $\varphi$ are described by the function $A(\varphi) := A[\psi(\varphi)]$. For such theories the two functions $A(\varphi)$ and $V(\varphi)$ completely define the low-energy predictions. Brans-Dicke theories \cite{BD} comprise the widely studied special case where $V(\varphi) = 0$ and $A(\varphi) = e^{a \, \varphi}$, for constant $a$.
In what follows we require no assumptions about the scalar-matter couplings, but our interest is in situations where the scalar potential, $V(\varphi) = \left. A^4(\psi) \, U(\psi) \right|_{\psi(\varphi)}$, is negligible (making $\varphi$ a minimally coupled massless scalar). The vacuum field equations obtained from (\ref{action}) when $V \simeq 0$ are very simple
\begin{eqnarray}
\label{fieldeq_tens}
\mathcal{R}_{\mu \nu}+ 2 \, \partial_{\mu} \varphi \, \partial_{\nu} \varphi &=& 0 \,, \\
\label{fieldeq_sc}
\Box \, \varphi = \frac{1}{\sqrt{-g}} \, \partial_\mu
\left( \sqrt{-g} \; g^{\mu\nu} \partial_\nu \varphi
\right) &=& 0 \,.
\end{eqnarray}
Time-independent, spherically symmetric solutions to these equations are known \cite{spheresoln, damourrev, BMQ}, and given by
\begin{equation}
{\hbox{d}} s^2 = - f(r) \, {\hbox{d}} t^2 + \frac{{\hbox{d}} r^2}{f(r)}
+ g(r) \, \Bigl( {\hbox{d}} \theta^2 + \sin^2 \theta \,
{\hbox{d}} \phi^2 \Bigr)
\quad \hbox{and} \quad
\varphi = \varphi(r) \,,
\end{equation}
where
\begin{eqnarray} \label{sphsymsolns}
f(r) &=& \left( 1 - \frac{\ell}{r} \right)^\alpha \nonumber\\
g(r) &=& r^2 \left( 1 - \frac{\ell}{r} \right)^{1-\alpha} \\
\hbox{and} \quad
\varphi(r) &=& \varphi_\infty +
\frac{q}{2} \ln \left( 1 -
\frac{\ell}{r} \right) \,,\nonumber
\end{eqnarray}
and the constants $\alpha$ and $q$ must satisfy $\alpha^2 + q^2 = 1$.
The Schwarzschild black hole corresponds to the choice $\alpha = 1$ and $q = 0$, in which case $\varphi = \varphi_\infty$ is constant, and the constant $\ell$ is revealed as the Schwarzschild radius: $\ell = 2 GM$. In this context the no-hair theorems \cite{nohair} state that $\alpha = 1$ and $q = 0$ is the only physically allowed black-hole solution. Within the family defined by eqs.~\pref{sphsymsolns}, the problem with the more general solutions is that $\varphi$ always diverges at $r = \ell$ whenever $q \ne 0$. Since the field equation \pref{fieldeq_tens} shows that the curvature scalar, $\mathcal{R}$, also diverges there, $r = \ell$ is an honest-to-God curvature singularity, rather than simply marking the coordinate singularity associated with an event horizon.
\subsection{Masses and scalar hair}
\label{sec:massschair}
We next pause to define scalar hair more precisely, for use in later sections. One way to define the (ADM) mass \cite{ADM, BMQ}, $M$, and the `scalar charge' \cite{damourrev, Stell1, StellUS}, $Q$, is through the large-$r$ asymptotic form of the metric and the scalar field\footnote{Notice our conventions here for $Q$ follow those of ref.~\cite{StellUS}, and so differ by a sign from those of refs.~\cite{damourrev, Stell1}, who define $\varphi = \varphi_\infty + G\omega/r + \cdots$.}
\begin{equation}
g_{tt} = -1 + \frac{2 G M}{r} + \cdots
\quad \hbox{and} \quad
\varphi = \varphi_\infty - \frac{GQ}{r} + \cdots
\,.
\end{equation}
In terms of these definitions we see that the integration constants $\ell$ and $q$ can be traded for $M$ and $Q$, with
\begin{equation}
2GM = \alpha \, \ell \quad \hbox{and} \quad
GQ = \frac{q \ell}{2} \,.
\end{equation}
For later purposes it is worth elaborating a bit on the extent to which this definition of $Q$ provides a useful definition of `charge'. The main observation is that the scalar equation of motion, eq.~\pref{fieldeq_sc}, can be read as a conservation law: $\partial_\mu J^\mu = 0$, for $J_\mu := \sqrt{-g} \; \partial_\mu \varphi$. Although the quantity $Q$ is {\em not} the charge --- {\em i.e.} the spatial integral of $J^t$ --- for this conserved current, it is `conserved' in the sense that eq.~\pref{fieldeq_sc} implies that the combination $\sqrt{-g} \; g^{rr} \varphi'$ is independent of $r$ (with $\varphi' := \partial_r \varphi$), for any static spherically symmetric configuration. $Q$ as defined above is equivalent to this $r$-independent quantity because
\begin{equation}
\sqrt{-g} \; g^{rr} \varphi' = f \, g \, \varphi' \,
\sin\theta = \left( 1 - \frac{\ell}{r} \right) r^2 \varphi' \sin \theta = \frac{q \ell}{2} \; \sin \theta = GQ \, \sin \theta \,.
\end{equation}
Our interest in the rest of this paper is in applications computing the radiation rate and post-Newtonian corrections to orbits, for which it is only the centre-of-mass motion of the black hole that is physically relevant. Furthermore, all objects are located far enough from one another to allow their weak mutual interaction to be described by the far-field regime. In this case it is more useful to define mass and charge in terms of the effective point-particle action of the black hole which governs the evolution of its centre of mass \cite{PPAction, DamEFGW, NRCMEFT, NRCMEFTrev},
\begin{equation} \label{PPactiondef}
S_p := \int_\Gamma {\hbox{d}} \tau \,
\mathcal{L}_p[\varphi, g_{\mu\nu}, \partial_\alpha \varphi, R_{\alpha\beta\gamma\delta}, \dot z, \cdots]\,,
\end{equation}
where $\mathcal{L}_p$ is a scalar whose dependence on the background metric and $\varphi$, and their derivatives depends on the physical properties of the object of interest ({\em e.g.} planet, star, neutron star, black hole). This dependence is obtained (see below) by matching the physical predictions of this effective theory to the predictions of a more microscopic description of the object in question \cite{damourrev, NRCMEFT}. In eq.~\pref{PPactiondef} $\mathcal{L}_p$ is evaluated along the point-particle world-line, $z^\mu(\tau)$, denoted as $\Gamma$, with $\tau$ representing proper time (arc-length) measured along $\Gamma$. ($\mathcal{L}_p$ could also depend on other variables for objects with internal degrees of freedom, like spin.) Over-dots denote the derivative with respect to $\tau$, as in $\dot z^\mu := {\hbox{d}} z^\mu/{\hbox{d}} \tau$.
This action has two important uses. First, it can be used as a source in the field equations for $\varphi$ and $g_{\mu\nu}$, and expressions for $\mathcal{L}_p$ can be chosen to make the resulting solutions agree with the far-field solutions computed from an explicit description of the relevant object. (For example, for a star these solutions can be compared to the far-field solutions obtained by matching to interior solutions to the equations of hydrostatic equilibrium inside the star.)
Second, its variation with respect to $z^\mu(\tau)$ can be used to determine the trajectory taken by the object moving within its local geometric environment, including any applied fields $\varphi$. For these latter applications it is useful to organize $\mathcal{L}_p$ into an expansion in derivatives of the background fields,
\begin{equation} \label{higherderivsv0}
S_p = \int_\Gamma {\hbox{d}} \tau \Bigl[ m(\varphi) + h(\varphi) \,
\dot z^\mu \dot z^\nu \, \partial_\mu \varphi
\, \partial_\nu \varphi + k(\varphi) \, \partial_\mu \varphi \, \partial^\mu \varphi + \cdots \Bigr] \,,
\end{equation}
since it is often only the first term, $\mathcal{L}_p \simeq m(\varphi)$ that is relevant for practical applications \cite{DamEFGW}. The coefficients like $m$, $h$ and $k$ take values whose order of magnitude can be determined on dimensional grounds in terms of the mass, $M$, size, $L$, and time-scale $T$ of the source in question; typically with $m \sim \mathcal{O}(M)$, and $k \sim \mathcal{O}(M L^2)$ and $h \sim \mathcal{O}(M T^2)$ and so on. (Here the characteristic time-scale is $T \sim L/v$, where $v$ denotes the typical speed of objects within the source; $v \simeq c = 1$ for relativistic sources like neutron stars.) For practical applications to post-Newtonian orbits it is usually the non-relativistic limit of the point-particle action evaluated in the field of the other body that is relevant, which we obtain by expanding in powers of $v$. (Although the systematics of this expansion are interesting in their own right \cite{DamEFGW,NRCMEFT,NRCMEFTrev} we do not follow these complications here.)
Our immediate interest is in constraining the function, $\mathcal{L}_p(\varphi)$, that describes the black-hole source that gives rise to the explicit scalar-tensor solutions given by eqs.~\pref{sphsymsolns}, above. This may be accomplished by solving the field equations in the presence of the point-particle action, $\delta (S_{g} + S_{p}) = 0$, for instance
\begin{equation} \label{scaleq_pp}
\partial_\mu \Bigl( \sqrt{-g} \; g^{\mu\nu} \partial_\nu
\varphi \Bigr) + 4\pi G \left( \frac{\delta \mathcal{L}_p}{\delta \varphi} \right) \; \delta^3[x - z(s)] = 0 \,,
\end{equation}
and comparing the far-field result to the solutions of eqs.~\pref{sphsymsolns}. We do not use the derivative expansion, eq.~\pref{higherderivsv0}, when doing so because this can break down when evaluating $S_p$ at self-field configurations sourced by the object itself, due to the divergence of these self-fields at the position of the source.\footnote{These divergences arise already at the classical level, and can be handled systematically by renormalizing the particle action \cite{PPAction, DamEFGW, NRCMEFT, NRCMEFTrev}, as is also relevant in other contexts \cite{BraneRenorm}.}
Taking a static source, and integrating eq.~\pref{scaleq_pp} at a fixed time over a spherical region centered on the source position implies in particular
\begin{equation} \label{QvsLp}
-G \, \left. \frac{\delta \mathcal{L}_p}{\delta \varphi}
\right|_p = \frac{1}{4\pi}
\oint_{r,t} {\hbox{d}} \theta {\hbox{d}} \phi \, \sqrt{-g} \; g^{rr} \varphi' = f(r) \, g(r) \, \varphi'(r)
= GQ \,,
\end{equation}
where the far-left-hand term is evaluated at the source position (which generically diverges and so must be renormalized). The Einstein equations give a similar expression involving the stress energy derived from $\mathcal{L}_p$. For example, for widely separated, weakly interacting objects moving within a field that asymptotes at large $r$ to a static constant configuration, $\varphi_\infty$, and assuming $\mathcal{L}_p \simeq m(\varphi)$, these considerations reproduce the standard results \cite{damourrev}
\begin{equation} \label{MQvsm}
M = m(\varphi_\infty)
\quad \hbox{and} \quad
Q = -m'(\varphi_\infty) \,.
\end{equation}
Notice that use of the weak, far-field expansion does {\em not} imply any assumptions about the nature of the underlying source, and in particular does not exclude it being a relativistic star or a black hole \cite{damourrev}. However, in the special case that the source {\em is} constructed purely from non-relativistic matter, and when this matter couples to $\varphi$ only through the Jordan-frame metric, $\hat g_{\mu\nu}$, we would additionally know that $m(\varphi) \, {\hbox{d}} \tau = m_0 \, {\hbox{d}} \hat \tau$, where $\hat \tau$ is measured using the metric $\hat g_{\mu\nu}$. In this case
\begin{equation}
m_{\scriptscriptstyle NR}(\varphi)
\simeq m_0 \, A(\varphi) \,,
\end{equation}
leading to $M_{\scriptscriptstyle NR} \simeq m_0 \, A(\varphi_\infty)$ and
\begin{equation} \label{NRQMrel}
Q_{\scriptscriptstyle NR} \simeq -m_0 A'(\varphi_\infty) = -M_{\scriptscriptstyle NR} \, a(\varphi_\infty) \,,
\end{equation}
where $a(\varphi) := A'(\varphi)/A(\varphi)$ \cite{damourrev}.
Eq.~\pref{NRQMrel} agrees with more detailed studies of stellar structure in scalar-tensor gravity \cite{Stell1,StellUS}, which also show how $Q/M$ differs from $a(\varphi_\infty)$ as the underlying stellar equation of state becomes more relativistic.
What is the function, $m(\varphi)$, appropriate when the object in question is a nonsingular, static black hole? From solutions eqs.~\pref{sphsymsolns} we see that static black hole solutions have a constant scalar profile, $\varphi = \varphi_\infty$, on which the asymptotic metric does not depend. Consequently $M = m(\varphi_\infty)$ is independent of $\varphi_\infty$ --- which is consistent with $Q = -m'(\varphi_\infty) = 0$, as found above.
For black holes the conclusion that $m(\varphi)$ must be independent of $\varphi$ should be very general, and should also apply to the coefficients $h(\varphi)$ and $k(\varphi)$ of the expansion, eq.~\pref{higherderivsv0}. It is this general because it is not necessary to couple to matter at all in order to form a black hole, and so the effective action for a black hole should be invariant under the shift symmetry, $\varphi \to \varphi + \hbox{constant}$, of the classical field equations, eqs.~\pref{fieldeq_tens} and \pref{fieldeq_sc}.
\subsection{Jacobson's Miracle Hair-Growth Formula}
\label{sec:bh}
For massless scalars the asymptotic value, $\varphi_\infty$, of the scalar field can be chosen freely, and is not related to the physics of a star or black hole. This is easily understood because far from the black hole the geometry becomes flat, and any constant value $\varphi = \varphi_\infty$ solves the scalar field equation, $\Box \,\varphi = 0$. However a function linear in time, $\varphi = \mu t$, is another solution to $\Box \, \varphi = 0$ for flat spacetime, and so one might ask what the black hole solution might be that asymptotes to this for large $r$.
This question need not be completely academic. For instance, for cosmological applications the scalar potential, $V(\varphi)$, need not precisely vanish and might drive a time evolution for $\varphi$ far from the black hole. For small $V$ this evolution could be adiabatic on the time-scales relevant to black-hole physics, and so be well-approximated by a slow, approximately linear, variation of $\varphi$ at infinity.
Although the exact solution to this problem is not known, a perturbative solution to eqs.~(\ref{fieldeq_tens}) and (\ref{fieldeq_sc}) is known with the property that $\varphi \to \varphi_\infty + \mu t$ as $r \to \infty$ \cite{jacobson}. This solution is given as an expansion about the black-hole solution,
\begin{eqnarray}
g_{\mu \nu} &=& g_{\mu \nu}^{(0)} + \epsilon^{2} g_{\mu \nu}^{(2)} + \ldots \,,
\nonumber
\\
\label{pertexp}
\varphi &=& \varphi^{(0)} +
\epsilon \varphi^{(1)} + \epsilon^{2} \varphi^{(2)} + \ldots \,,
\end{eqnarray}
where $\epsilon$ is a small dimensionless parameter and
\begin{equation}
\label{schwarzschild}
g_{\mu \nu}^{(0)} {\hbox{d}} x^{\mu} {\hbox{d}} x^{\nu} =
- \left( 1 - \frac{2GM}{r} \right)
\, {\hbox{d}} t^{2} + \frac{{\hbox{d}} r^{2}}{1-2GM/r} + r^{2} {\hbox{d}}\Omega_{2}^{2}
\end{equation}
is the Schwarzschild metric. $\varphi^{(0)} = \varphi_\infty$ is a constant (which we can take to vanish with no loss of generality), and
\begin{equation}
\label{jacobson_sol}
\varphi^{(1)}_{\pm} = \frac{t}{2GM} \pm \log \left|
1 - \frac{2GM}{r} \right|
\end{equation}
is the leading scalar-field perturbation. Both choices of sign solve eq.~(\ref{fieldeq_sc}), but we will shortly see that only $\varphi_{+}$ is physically realistic for our later applications. The back-reaction of the metric first arises at $\mathcal{O}(\epsilon^{2})$, because equation (\ref{fieldeq_tens}) is quadratic in $\partial \varphi$.
Comparing this solution at large $r$ to $\mu t$ shows that $\mu = \epsilon/2GM$ and so
\begin{equation}
\varphi_{\pm} = \mu t \pm 2GM\mu \log \left|
1-\frac{2GM}{r}\right| +
\mathcal{O}[(GM\mu)^{2}] \,.
\end{equation}
Thus an expansion in $\epsilon \ll 1$ corresponds to choosing the time scale $\mu^{-1}$ over which the asymptotic scalar field $\varphi$ evolves to be very long relative to the light-crossing time scale, $2GM/c^{3}$, of the black hole.
The necessity for the $r$-dependence in $\varphi^{(1)}$ can be seen by asking how the solution behaves near the black-hole event horizon. In ref.~\cite{jacobson} Jacobson shows that the singularity at $r = 2GM$ in the spatial part of $\varphi^{(1)}$ is required to cancel the singularity due to the breakdown of the coordinate $t$ at the horizon. To see this, we re-write it using coordinates that do not break down there, such as
\begin{equation}
u = t - r_{\star} \,, \qquad
v = t + r_{\star} \,,
\end{equation}
where
\begin{equation}
r_{\star} = r + 2GM \log \left| \frac{r}{2GM} - 1 \right|
\end{equation}
is the `tortoise' coordinate. Trading $t$ for $v$, the Schwarzschild metric is
\begin{equation}
g_{\mu \nu}^{(0)} {\hbox{d}} x^{\mu} {\hbox{d}} x^{\nu}
= -\left( 1 - \frac{2GM}{r}\right) {\hbox{d}} v^2
+ 2 \, {\hbox{d}} r \,
{\hbox{d}} v + r^{2} {\hbox{d}}\Omega^{2}_2 \,,
\end{equation}
and the linearized scalar solutions become
\begin{equation}
\varphi^{(1)}_{+} = \left( \frac{v-r}{2GM} \right)
- \log \left(
\frac{r}{2GM} \right) \,,
\quad
\varphi^{(1)}_{-} = \left( \frac{u+r}{2GM}
\right) + \log\left( \frac{r}{2GM} \right) \,.
\end{equation}
Now, for an eternal black hole the future event horizon is at $t \to \infty$ and $r=2GM$, which corresponds to $u \to \infty$ and finite $v$. The past horizon is at $t\to -\infty$ and $r=2GM$, which corresponds to $v\to -\infty$ and finite $u$. Therefore $\varphi^{(1)}_{+}$ is regular at the future horizon and singular at the past horizon, while $\varphi^{(1)}_{-}$ is regular at the past horizon and singular at the future horizon. Thus, it is the perturbation $\varphi_{+}^{(1)}$ that is relevant for a physically realistic black hole formed by gravitational collapse with no past horizon.
\subsubsection*{Quantifying how much hair}
Because $\varphi^{(1)}_+$ has a radial nontrivial profile, it also introduces a nonzero scalar charge. Expanding eq.~(\ref{jacobson_sol}) in powers of $1/r$ yields
\begin{equation}
\varphi^{(1)}_{+} = \frac{t}{2GM} - \frac{2GM}{r}
+ \mathcal{O}\left( \frac{1}{r^{2}} \right) \,,
\end{equation}
and so comparing the $1/r$ term with $-GQ/r$ implies
\begin{equation}
\frac{Q}{M} = 2\, \epsilon = 4GM \mu \,,
\end{equation}
indicating that asymptotic scalar time-dependence gives black holes nonzero scalar charge.
Another route to the same conclusion is to recognize that so long as we restrict to spherically symmetric field configurations with a time-independent metric and $\partial_t \varphi$ nonzero but $t$-independent, it follows that $\Box \, \varphi = 0$ still implies the $r$-independence of $\sqrt{-g} \; g^{r\mu} \partial_\mu \varphi = f \, g \, \varphi' \, \sin \theta$. Using Jacobson's asymptotic expression for large $r$ then shows its value for all nonzero $r$ must be
\begin{equation}
f(r) \, g(r) \, \varphi' = \lim_{r \to \infty} f(r) \, g(r) \, \varphi' = \lim_{r \to \infty} r^2 \varphi' = 2 GM \, \epsilon \,,
\end{equation}
and so $GQ = (2 GM)^2 \mu$, as concluded above. What is important for later post-Newtonian applications is the black hole's dimensionless charge-to-mass ratio,
\begin{equation}
\label{efscmcoupl}
a := \frac{Q}{M} = 2 \epsilon = 4 GM \mu \,.
\end{equation}
We saw above that for ordinary stars with non-relativistic equations of state in quasi Brans-Dicke theories this quantity would play the role of the effective coupling, $a(\varphi_\infty)$, to the scalar field far from the star. This need no longer be true for relativistic systems like neutron stars or black holes \cite{damourrev,Stell1,StellUS}.
The utility of this last derivation lies in its connection to eq.~\pref{QvsLp}, which shows that the point-particle effective action for such a black hole must satisfy
\begin{equation}
-G \, \left. \frac{\delta \mathcal{L}_p}{\delta \varphi}
\right|_p = f(r) \, g(r) \, \varphi'(r)
= (2 GM)^2 \mu \,.
\end{equation}
Notice that this conclusion is not inconsistent with the black-hole shift symmetry, since this symmetry does not preclude a nonzero functional derivative for $S_p$.
\subsubsection*{Orbital corrections}
Where the existence of the shift symmetry {\em does} play an important role is in how the black hole responds to applied scalar and gravitational fields (such as those due to a companion within a binary system). For these applications a derivative expansion is appropriate, and gives
\begin{equation} \label{higherderivs}
S_p = \int_\Gamma {\hbox{d}} \tau \Bigl[ m(\varphi) + h(\varphi) \,
\dot z^\mu \dot z^\nu \, \partial_\mu \varphi
\, \partial_\nu \varphi + k(\varphi) \, \partial_\mu \varphi \, \partial^\mu \varphi + \cdots \Bigr] \,,
\end{equation}
where the functions $h(\varphi)$ and $k(\varphi)$ are to be determined by an appropriate matching calculation. A possible term linear in $\dot z^\mu$ --- {\em i.e.} $\int_\Gamma {\hbox{d}} \tau \, f(\varphi) \, \dot z^\mu \, \partial_\mu \varphi$ --- is not written in eq.~\pref{higherderivs} because it is a total derivative and so does not contribute to the local equations of motion, and several other terms involving the background Ricci tensor can be removed by performing a judicious field redefinition \cite{DamEFGW, GREFT}. The ellipses include terms with either more powers of $\partial_\mu \varphi$, derivatives of the metric, or higher derivatives of $\varphi$.
As before, for applications to black holes the functions $m(\varphi)$, $h(\varphi)$ and $k(\varphi)$ are independent of $\varphi$, since black hole solutions do not require couplings between $\varphi$ and matter and so do not break the classical symmetry under constant shifts: $\varphi \to \varphi + \omega$. When calculating the response of the black hole to fluctuations, $\delta \varphi = \varphi - \varphi_\infty - \mu t$, about the asymptotic scalar, the above action shows the leading terms linear in $\delta \varphi$ are
\begin{eqnarray} \label{NewSp}
\delta S_p &\simeq& 2\mu \int_\Gamma {\hbox{d}} \tau \, \Bigl[
\Bigl( h \, \dot t^2 + k \, g^{tt} \Bigr)
\delta \varphi_{,\,t} + h \, \dot t \, \dot r \,
\delta \varphi_{,\,r} \Bigr] \nonumber\\
&\simeq& 2\mu \int_\Gamma {\hbox{d}} \tau \, \Bigl[
(h - k) \, \delta \varphi_{,\,t} + h \, \dot r \,
\delta \varphi_{,\,r} + \mathcal{O}(v^2) \Bigr] \,,
\end{eqnarray}
where $\delta \varphi_{,\,r} := \partial_r \delta \varphi$ and $\delta \varphi_{,\,t} := \partial_t \delta \varphi$. Notice that the shift symmetry, $\varphi \to \varphi + \omega$, guarantees that $\delta \varphi$ always appears differentiated, and so in principle eq.~\pref{NewSp} gives a different response than would the expansion $m(\varphi) \simeq m_0 + m'_0 \, \delta \varphi + \cdots$ that dominates the response of other matter sources, like neutron stars.
Considerations like these show that computing the changes to black hole orbits due to an asymptotic scalar evolution can be complicated, even in the post-Newtonian approximation. However the same issues do not similarly complicate the rate that scalars are emitted by black holes in binary systems, provided we work to leading nontrivial order in $\mu$. This is because the black hole scalar charge is itself of order $\mu$, and so the leading contribution to dipole radiation comes from the radiation due to a charge of order $\mu$ moving along the zeroth-order Schwarzschild orbits. Although the $\mathcal{O}(\mu^2)$ corrections to these orbits are complicated to compute, they are subdominant contributions to the scalar dipole radiation.
\subsection{Dipole radiation from binaries}
\label{sec:bhbinfr}
Consider first the problem of how a single black hole radiates energy into the scalar field as a function of its trajectory, which we assume (temporarily) to be given. Although we assume in this section that the scalar is massless, it really suffices that it be light enough to have modes into which the system can radiate. Since periodic motion with angular frequency $\Omega$ typically radiates dominantly into modes with angular frequency $\Omega$, this requires $m c^2/\hbar \ll \Omega$ \cite{cloutier}.
\subsubsection*{Scalar radiation from a single particle moving on a given trajectory}
In order to find the power emitted into scalar radiation, the scalar field
far from the source is written as
\begin{equation}
\varphi({\bf x}, t) = \mu t - \frac{G}{r
} \left[
\psi + \frac{x^{i}}{r
} \,
\psi_i' +
\frac{x^{i}x^{j}}{2r^{2}
} \;
\psi_{ij}'' + \ldots
\right]
+ \mathcal{O}\left(\frac{1}{r^{2}}\right) \,,
\end{equation}
where $\psi_{i_{1}\cdots i_{l}}(t)$ is the time-dependent STF (symmetric trace-free) multipole moment of
order $l$ \cite{multipole}.
Primes denote differentiation with respect to time and the right-hand side is to be evaluated at the retarded time, $t_r := t - r$, where $r := |{\bf x}|$ denotes the (large) distance from the source to the field point of interest.
The terms in the square brackets may be thought of as a series in $v$, the typical velocity
of the source. Terms that fall off as $1/r^{2}$ and faster do not contribute to the emitted power.
The contribution to the emitted power from the dipole moment $\psi_{i}$ is given by \cite{damourrev}
\begin{equation}
\label{dippwr1}
F_{\varphi}^{\rm dip} = \frac{G}{3
} \sum_{i}
\left( {\psi_i''}
\right)^{2} \,.
\end{equation}
Explicitly, in terms of the non-relativistic source density, $\varrho({\bf x}, t)$, defined by the field equation
\begin{equation}
\label{scsource}
\Box_{f} \varphi = 4 \pi G \varrho
+ \mathcal{O}(v^{4
) \,,
\end{equation}
where $v$ is the typical velocity of the source, and $\Box_{f}$ is the flat-space d'Alembertian operator, we have
\begin{eqnarray} \label{dipmom}
\psi(t_r) &=& \int {\hbox{d}}^{3}x
\; \varrho(t_r, {\bf x}) +
\mathcal{O}(v^{2
) \nonumber\\
\psi_{i}(t_r) &=& \int {\hbox{d}}^{3}x
\; \varrho(t_r, {\bf x}) \, x_{i} +
\mathcal{O}(v^{2
) \,,
\end{eqnarray}
and so on.
For a point object with scalar charge $Q$ and trajectory ${\bf z}(t)$ we have $\varrho(t, {\bf x}) = Q \, \delta^{3}[{\bf x} - {\bf z}(t) ]$, and so $\psi = Q + \mathcal{O}(v^2)$ and $\psi_{i}(t) = Q \, z_i(t) + \mathcal{O}(v^2)$, {\em etc}. Equations (\ref{dippwr1}), (\ref{scsource}) and (\ref{dipmom}) show that the $Q$ that appears in standard calculations of the scalar dipole radiation rate is the same $Q$ that is identified in section \ref{sec:bh} using Jacobson's Miracle Hair Growth formula for asymptotically varying scalar fields. This allows the use of standard expressions for scalar radiation when computing the leading dipole scalar radiation rate for binary systems.
\subsubsection*{Scalar radiation from binary systems}
Next consider the radiation rate from a pair of compact objects having masses $M_{{\scriptscriptstyle A}}, M_{{\scriptscriptstyle B}}$, and scalar charges, $Q_{\scriptscriptstyle A}$, $Q_{\scriptscriptstyle B}$. An important role is played by the ratios $a_{{\scriptscriptstyle A}} = Q_{\scriptscriptstyle A}/M_{\scriptscriptstyle A}, a_{{\scriptscriptstyle B}} = Q_{\scriptscriptstyle B}/M_{\scriptscriptstyle B}$. As emphasized earlier, this ratio would be the asymptotic scalar-matter coupling, $a(\varphi_\infty)$, for ordinary stars in Brans-Dicke-like theories, but this need not be true for objects with relativistic structure, like neutron stars or black holes.
As usual, when all other things are equal, we expect scalar radiation to be well constrained by binary orbits because it can arise as dipole radiation. This is by contrast with gravitational radiation, which first appears at quadrupole order and so with a rate that is suppressed relative to dipole radiation by additional factors of $v$, with $v$ being a representative orbital speed (which for numerical purposes we take to be the average orbital speed).
To quantify this, let $F^{\rm mon}$, $F^{\rm dip}$ and $F^{\rm quad}$ respectively denote the power in monopole, dipole and quadrupole radiation, integrated over an orbital period. For objects whose orbits are close to Newtonian the total energy loss is the sum of tensor and scalar contributions, $F = F_g + F_\varphi$, with the leading gravitational contribution being quadrupole radiation, $F_g \simeq F_g^{\rm quad} + \cdots$. By contrast, for scalars monopole and dipole radiation is possible, $F_\varphi \simeq F_\varphi^{\rm mon} + F_\varphi^{\rm dip} + F_\varphi^{\rm quad} + \cdots$.
In order of magnitude $F_{\varphi}^{\rm mon} = \mathcal{O}[a_{{\scriptscriptstyle A,B}}^{2} v^{5}]$, $F_{\varphi}^{\rm dip} = \mathcal{O}[a_{{\scriptscriptstyle A,B}}^{2} v^{3}]$, $F^{\rm quad}_{g} = \mathcal{O}[v^{5}]$ and $F^{\rm quad}_{\varphi} = \mathcal{O}[a_{{\scriptscriptstyle A,B}}^{2} v^{5}]$. (Notice that $F_{\varphi}^{\rm mon}$ is order $v^{5}$ rather than order $v$ because the scalar charges $Q_{{\scriptscriptstyle A,B}}$ are time-independent.)
Constraints on scalar-tensor theories arise as upper bounds on the scalar radiation rate for systems where the gravitational rate is observed. To this end define the ratio
\begin{equation}
\xi := \frac{F_\varphi}{F_g} \simeq \frac{F^{\rm dip}_\varphi}{F^{\rm quad}_g} = \mathcal{O}\left[ \frac{a_{{\scriptscriptstyle A,B}}^{2}}{v^2} \right] \,.
\end{equation}
Explicit formulae for the orbitally averaged power emitted into scalar and gravity waves are calculated in chapter 6 of \cite{damourrev}, which finds (re-introducing the factors of $c$)
\begin{eqnarray}
\label{dippwr}
F^{\rm dip}_\varphi &=& \frac{\nu^{2} (GM\Omega)^{8/3} (1+e^{2}/2)(a_{{\scriptscriptstyle A}}-a_{{\scriptscriptstyle B}})^{2}}
{3c^{3}G(1-e^{2})^{5/2}} \Bigl(1 + \mathcal{O}[(v/c)^{2}]
\Bigr) \,,
\\
\label{quadpwr}
F^{\rm quad}_{g} &=& \frac{32 \nu^{2} (GM\Omega)^{10/3}(1+73e^{2}/24 + 37e^{4}/96)}
{5c^{5}G(1-e^{2})^{7/2}} \Bigl(1 + \mathcal{O}[(v/c)^{2}] \Bigr) \,,
\end{eqnarray}
where $e$ is the orbit eccentricity, $\Omega = 2\pi/P$ where $P$ is the orbital period, $M=M_{{\scriptscriptstyle A}}+M_{{\scriptscriptstyle B}}$ is the total mass, and $\nu = M_{{\scriptscriptstyle A}}M_{{\scriptscriptstyle B}}/M^{2}$. Using these, the ratio $\xi$ becomes
\begin{eqnarray}
\label{qest1}
\xi &=& \frac{5c^{2}(a_{{\scriptscriptstyle A}}-a_{{\scriptscriptstyle B}})^{2}(1-e^{2})(1+e^{2}/2)}
{96(GM\Omega)^{2/3}(1+73e^{2}/24+37e^{4}/96)} \Bigl(1 + \mathcal{O}[(v/c)^{2}] \Bigr)
\nonumber
\\
\label{qest}
&\simeq& \frac{5}{96}
\, \frac{
(a_{{\scriptscriptstyle A}}-a_{{\scriptscriptstyle B}})^{2}}{(v/c)^2} \; f(e)
\Bigl( 1 + \mathcal{O}[(v/c)^{2}] \Bigr) \,.
\end{eqnarray}
Notice that $(GM\Omega/c^{3})^{2/3} = GM/\frak{a} c^2 = (v/c)^{2}$ in the Keplerian limit, where $\frak{a}$ is the semi-major axis of the orbit and $v$ is the average orbital speed.
For two black holes with an asymptotically static scalar field we have $a_{\scriptscriptstyle A} = a_{\scriptscriptstyle B} = 0$ because of the no-hair condition $Q_{\scriptscriptstyle A} = Q_{\scriptscriptstyle B} = 0$, as is well-known to hold for Kerr and Schwarzschild black holes in general relativity \cite{willzag}. This implies in particular the usual absence of dipole radiation, $\xi \simeq 0$, as expected.
When the asymptotic scalar varies slowly in time --- {\em i.e.} $\mu \ne 0$ --- we've seen that the above expressions can also be used to compute the leading contribution to the scalar dipole radiation, with the scalar charge given as in section \ref{sec:bh}: $a_{\scriptscriptstyle A} = 2 \epsilon_{\scriptscriptstyle A}$ and $a_{\scriptscriptstyle B} = 2 \epsilon_{\scriptscriptstyle B}$. Since both black holes see the same asymptotic scalar field far from the binary system, consistency requires
\begin{equation}
\frac{\epsilon_{{\scriptscriptstyle A}}}{2GM_{{\scriptscriptstyle A}}} =
\frac{\epsilon_{{\scriptscriptstyle B}}}{2GM_{{\scriptscriptstyle B}}} = \mu \,,
\end{equation}
and so because the two black holes share the same asymptotic time-dependent scalar field, their scalar charges are related to one another.
The quantity appearing in $\xi$ is $a_{\scriptscriptstyle A} - a_{\scriptscriptstyle B} = 2(\epsilon_{\scriptscriptstyle A} - \epsilon_{\scriptscriptstyle B}) = 4 G \mu (M_{\scriptscriptstyle A} - M_{\scriptscriptstyle B})$. We see that inspiralling black holes with differing masses will radiate dipole radiation when situated within a slowly varying asymptotic scalar field, provided only that the scalar is light enough to receive the radiation. The amount of radiation is guaranteed to be small by virtue of the approximation $\epsilon_{{\scriptscriptstyle A,B}} \ll 1$ that underlies our linearized analysis.
This calculation assumes $Q_{\scriptscriptstyle A}$ and $M_{\scriptscriptstyle A}$ are fixed quantities over any one orbit, and this is true in the present case if $\epsilon_{\scriptscriptstyle A,B} \ll 1$ and $\mu$ doesn't change in time. Notice in particular that (at leading order) we need {\em not} assume $\mu \ll \Omega$ for the radiation-loss calculation to be valid.
\section{The Quasar OJ287}
\label{sec:oj287}
We next apply the formalism of section \ref{sec:bhbinfr} to OJ287. This is a quasar situated at redshift $z=0.306$ that undergoes periodic bursts. This system has been argued to consist of two gravitationally-bound black holes \cite{oj287, oj287GR, oj287nohair}, with the bursts occurring when the `small' black hole (black hole B, with mass $M_{\scriptscriptstyle B} \simeq 10^8$ $M_\odot$) passes through the accretion disc of the larger one (black hole A, which is a monster with $M_{\scriptscriptstyle A} \simeq 10^{10}$ $M_\odot$).
The orbital parameters of the system have been reconstructed from the observed bursting pattern, leading to the orbital parameters summarized in table \ref{orbpars}. In particular, the success of the orbital description requires the orbital decay due to the emission of gravitational radiation, with the observed radiated power being within $6\%$ of the prediction of General Relativity \cite{oj287GR}.
Taking this binary black hole interpretation as correct, we may use its success to infer an upper limit for the time-variation of any ambient light scalar fields that were present in the black hole neighborhood at an epoch $z = 0.306$. Notice that this does {\em not} require making any assumptions about the nature of the scalar-matter couplings. This bound applies to any scalar fields at this epoch, provided that they are light enough that it is possible to radiate into their scalar modes. This requires the mass of the scalar to be much lighter than the angular frequency, $\Omega$. Given that the orbital period is about 9 years, the corresponding scalar mass limit becomes $m \ll 10^{-23}$ eV/$c^2$ \cite{cloutier}.
Using the black hole orbital parameters in equation (\ref{qest}), we find $\xi \sim 0.94 (a_{{\scriptscriptstyle A}}-a_{{\scriptscriptstyle B}})^{2}$, and so imposing $\xi < 0.06$ gives the constraint $|a_{{\scriptscriptstyle A}}-a_{{\scriptscriptstyle B}}| < 0.25$ and so $|\epsilon_{{\scriptscriptstyle A}}-\epsilon_{{\scriptscriptstyle B}}| < 0.13$. We have $\epsilon_{{\scriptscriptstyle A,B}} = 2GM_{{\scriptscriptstyle A,B}} \, \mu/c^{3}$, where $2GM_{{\scriptscriptstyle A}}/c^{3} \sim 1.8 \cdot 10^{5}\ {\rm s} \sim 2\ {\rm day}$ is the light-crossing time of black hole A, and $2GM_{{\scriptscriptstyle B}}/c^{3} \sim 1.4 \cdot 10^{3}\ {\rm s} \sim 23\ {\rm min}$ is that of black hole B.
This finally implies $|\mu^{-1}| > 1.4 \cdot 10^{6}\ {\rm s} \sim 16\ {\rm days}$. This limit holds even though it is much smaller than the orbital period, because the instantaneous radiation rate relies only on the constancy of $\mu$ and the black hole masses and scalar charges. Our upper bound on $|\mu|$ implies $|\epsilon_{{\scriptscriptstyle A}}| < 0.13$ and $|\epsilon_{{\scriptscriptstyle B}}| < 1.0 \cdot 10^{-3}$. This is consistent with our linearization assumption $\epsilon_{{\scriptscriptstyle A,B}} \ll 1$.
\begin{table}
\begin{tabular}{|l|l|l|}
\hline
Parameter & Meaning & Value \\
\hline
$\dot \omega$ & Precession rate & $39.1(1)^{\circ}$ per orbit \\
$M_{{\scriptscriptstyle A}}$ & Mass of black hole A & $1.84(1) \cdot 10^{10} M_{\odot}$ \\
$M_{{\scriptscriptstyle B}}$ & Mass of black hole B & $1.40(3) \cdot 10^{8} M_{\odot}$ \\
$\chi_{{\scriptscriptstyle A}}$ & Dimensionless Kerr parameter of black hole A & $0.28(3)$ \\
$e$ & Orbital eccentricity & $0.658(1)$ \\
$P$ & Orbital period & $\sim 9\ {\rm years} \sim 2.8 \cdot 10^{8} {\rm s}$ \\
$\frak{a}$ & Semi-major axis & $\sim 1.7 \cdot 10^{15} {\rm m}$ \\
$v_{\rm orb}$ & Typical orbital velocity & $\sim 0.13c$ \\
\hline
\end{tabular}
\caption{Orbital parameters of OJ287.
These are the `intrinsic' values, as measured in the centre-of-mass of the binary. The values measured on Earth differ by a redshift factor. Data taken from \cite{oj287pars}. Uncertainties are given at the 3 sigma level.}
\label{orbpars}
\end{table}
Notice also that the spin of black hole $A$ is claimed to have been measured \cite{oj287spin}. The solution described in section \ref{sec:bh} can be generalized to rotating black holes \cite{jacobson}, and the analogue of equation
(\ref{efscmcoupl}) for Kerr black holes is
\begin{equation}
\label{scchkerr}
a_{{\scriptscriptstyle A}} = 2GM_{{\scriptscriptstyle A}} \mu (1 + \sqrt{1-\chi_{{\scriptscriptstyle A}}^{2}}) \,,
\end{equation}
where $\chi_{{\scriptscriptstyle A}} = cJ_{{\scriptscriptstyle A}}/GM_{{\scriptscriptstyle A}}^{2}$ is the dimensionless Kerr
parameter, and $J_{{\scriptscriptstyle A}}$ is the angular momentum of the black hole.
It follows from equation (\ref{scchkerr}) that taking spin into account
does not significantly change our bounds.
\section{Summary and Outlook}
\label{sec:summary}
In this paper we use Jacobson's Miracle Hair-Growth Formula to show that inspiralling black holes can radiate dipole scalar radiation whenever there exists a scalar whose mass is small enough to be radiated (typically this requires the mass to be much smaller than the orbital period, $m \ll \Omega$). We then apply this radiation calculation to the quasar OJ287, which has been interpreted as a very massive double black hole binary. Using the inferred black hole masses and orbital properties, and the success of the GR prediction for the orbital decay rate, we conclude that any ambient scalar fields lighter than $m < 10^{-23}$ eV at this epoch (redshift $z = 0.306$) must have instantaneous time evolution over scales larger than $|\mu^{-1}| \gtrsim 16$ days.
Although this is not a spectacular bound compared with cosmological time evolution, there are two things about it that are quite remarkable. First, it is completely independent of how the light scalar couples to matter since this does not enter into the amount of acquired black hole hair. Second, it directly constrains the instantaneous scalar evolution rate, rather than an average change in scalar field values over cosmological time intervals.
This system is representative of the rich kinds of physics that can arise for gravity/scalar systems with time-dependent environments, and indicates the wealth of information that is likely to become available once gravitational radiation is detected from astrophysical sources.
Direct detection of gravitational waves in the future will provide a systematic means of measuring the properties of double black hole binaries, making it possible to map out the bounds on $|\mu^{-1}|$ at different redshifts. In preparation for these happy times it would be useful to predict in more detail the response of these systems to scalar-tensor theories, and in particular how their orbits respond to their scalar `environment'.
Alternatively, if the scalars are not precisely massless they will propagate with different speeds than those of the primary gravitational signal, allowing potential time delays in gravitational wave signals whose presence could signal the existence of a new scalar component to gravity.
\section*{Acknowledgements}
We thank Craig Heinke, Ted Jacoboson, Laura Parker, Ethan Vishniac and James Wadsley for helpful discussions. Our research was supported in part by funds from the Natural Sciences and Engineering Research Council (NSERC) of Canada. Research at the Perimeter Institute is supported in part by the Government of Canada through Industry Canada, and by the Province of Ontario through the Ministry of Research and Information (MRI).
|
\section{Introduction}
A seemingly straightforward way to save energy of a mobile device is
to offload work to a more powerful machine. Several frameworks have
been proposed for \textit{computation offloading} including
MAUI\cite{cuervo10maui}, Cuckoo\cite{kemp10cuckoo},
CloneCloud\cite{chun11clonecloud}, and
ThinkAir\cite{sokol11thinkair}. They support method-level migration of
software execution and try to trade the energy spent in migration for
the savings gained from the reduction of local computation.
These frameworks share two common characteristics. One is that they fail to provide any tools or guidelines to help offload existing programs. The other is that they
focus only on heavy computation offloading. We challenge these design
decisions on the basis of our findings and argue that an alternative path should be taken for two reasons. First, automated offloading using existing frameworks has
certain limitations. To make existing applications offloadable,
modifications to source code is often needed and the process is
laborious. Second, many of the popular mobile applications, such as
Facebook and Twitter, require almost no computation, but a lot of
communication. Intuitively, offloading traffic seems to make no sense since content needs to reach the mobile device anyway. However, energy savings gained through \textit{communication
offloading} are possible because communication often contains costly signaling traffic \cite{signals103g}, part of which can be suppressed. Furthermore, packet
interval patterns and throughput, both of which have
a significant impact on communication cost\cite{yu10eenergy}, can be optimized by means of offloading.
To address these issues, we propose a toolkit, namely SmartDiet, which helps
application developers to study the offloadability of existing applications and, in turn, to implement offloading in an energy-efficient
manner. The toolkit identifies potential trouble spots in Android applications' Java source
code. It also estimates current energy usage at a fine-grained level in order to provide estimates of how much could be saved by offloading, and analyzes the software structure to identify opportunities for further savings. The following three points summarize our contributions.
1) We take the first look into the feasibility of applying
communication offloading to mobile applications, and analyze the
factors that may limit the energy savings. We use an open source
Twitter client to exemplify the associated issues.
2) We propose a toolkit, SmartDiet, for identifying constraints from existing program code. It guides programmers to improve the design and implementation of programs so that the existing offloading frameworks can be better utilized.
3) SmartDiet provides a novel way to estimate the savings in
communication energy cost. Our method is more fine-grained than those used in
current offloading frameworks, and, moreover, takes traffic patterns and
power saving modes into account. This part of the toolkit could also be
integrated into existing frameworks to enhance the accuracy of their
runtime cost estimation.
We motivate the need for SmartDiet in Section \ref{sec:use-case} by examining an offloading use case . We then describe our vision and current state of the toolkit in Section \ref{sec:toolkit}. Section \ref{sec:challenges} discusses the remaining challenges and future work before conclusions.
\section{Related work}
Two main approaches have been suggested for mobile application
offloading. MAUI\cite{cuervo10maui},
Cuckoo\cite{kemp10cuckoo} and ThinkAir\cite{sokol11thinkair} implement a framework on top of
the existing runtime system. These
three systems are fairly easy to deploy because they only require access to the program source code, and they do not need any special support from the operating system. The second approach, used by
CloneCloud\cite{chun11clonecloud}, is to modify the underlying virtual
machine or operating system in order to implement richer mechanisms for
offloading. CloneCloud is a fully automated system and does not
require having the source code of the program, because it
works directly on bytecode. We
claim that the developer should participate in the offloading
process and therefore focus on the first approach.
Specific solutions, such as Catnap\cite{dogar10catnap}, have been proposed in the literature for reducing the communication energy cost by
applying a proxy or middlebox approach. However, these solutions will provide energy savings only for the communication part of the
program, whereas offloading simultaneously provides savings in computational costs. Furthermore, since systems such as Catnap do not execute
application logic at the proxy, all traffic must eventually reach the mobile device. With smart offloading, some part of
the traffic (e.g., signaling) might never need to reach the
mobile device, because the offloaded part of the program handles it
directly.
\section{Use case: AndTweet}
\label{sec:use-case}
To gain experience in offloading existing network-intensive programs, we offloaded the communication part of a typical such application using ThinkAir. The application was AndTweet, an open source Android Twitter client which has been downloaded from the Android Market site over 5000 times. Over the course of this study, we had to overcome a number of obstracles in our efforts to remotely execute parts of the application. After resolving these issues, we measured the energy consumption of offloaded and non-modified versions of AndTweet. We show that, although somewhat promising, the results vary considerably depending on circumstances. Our experiences provide the motivation for our toolkit which we detail in the remaining sections.
\subsection{Offloading setup}
We used ThinkAir offloading system\cite{sokol11thinkair}, which
allows us to offload a selected set of methods to a remote server. We
used a Google Nexus One with Android 2.3 as the local device and a
virtual machine running the Android x86 port as the remote execution
platform. We chose the offloaded methods manually and disabled all the
dynamic decision making features of ThinkAir.
Upon starting an application, an execution controller from ThinkAir is
instantiated and an application image is sent to a remote
server. Whenever a method marked as offloadable is executed,
the controller transfers the execution to the server. When a method is executed remotely, the client serializes the class instance of the called method and all of its arguments using Java's
serialization APIs. The server de-serializes these objects,
invokes the specified method, and sends the returned value (or
exception) back to the client.
\subsection{Problems encountered during offloading AndTweet}
\label{sec:constraints-of-remote-execution}
We offloaded as many parts of AndTweet as possible which communicate with the Twitter backend. Our goal was to reduce the signaling traffic and hence save energy. We identified events in the user interface (UI) which trigger network requests and proceeded to search for methods whose execution could be migrated to the remote server, starting from the method which handles the UI event. We encountered several challenges, related to 1) methods accessing local hardware, 2) methods whose migration to remote server was not possible without modifications and 3) methods accessing state that is not correctly synchronized between the device and the server and therefore cause unexpected behavior.
A remotely executed method cannot interact with the UI or other hardware resources, because these resources only exist at the client. AndTweet mixes application logic and handling of UI in many of its methods that are interesting in terms of offloading. These pieces of application logic, which might be offloadable on their own, are tied to the device. To overcome the restrictions of UI interactions, we tried to offload the methods that UI handling methods directly invoke. Here we could find Twitter-specific abstractions, such as friend timeline, which is a list of timestamped messages from the people you follow in Twitter. This set of methods, however, contained another category of issues.
In order to migrate the execution of a method to a remote server, the offloading system must transfer the related dependencies over the network. Section \ref{sec:constraint-identification-tool} discusses migration further, but in case of ThinkAir this sets the requirement that the encapsulating class of an offloaded method must be serializable using Java's serialization APIs. In addition, they must not access any state outside the serialized context, because the changes are not automatically synchronized either to the remote server or back. The classes encapsulating Twitter timelines contained instances of non-serializable classes. AndTweet stores some of its internal state, for example cryptographical tokens required for authentication, using an Android standard library class SharedPreferences, which is not serializable. Similarly, HttpClient, the de-facto class for doing HTTP communication in an Android application, is also not serializable.
Methods using HttpClient were fixed by instantiating a new HttpClient in the remote server. Existing instances now render migration unnecessary to either direction. SharedPreferences, on the other hand, was more problematic, because behind the scenes it uses the local file system to save its state. Instantiating a new SharedPreferences in the remote server, and blindly using it would result in different states between the client and the server. Therefore, we needed to manually implement a mechanism which synchronizes the remote SharedPreferences before and after a method is executed remotely. Otherwise the cryptographical tokens were out-of-sync between the device and the remote execution server. The problems with state synchronization are difficult to resolve because in the worst case no errors are reported. Instead, the method just does something wrong and the program, not to mention the developer, is unaware about it.
Detecting and fixing the aforementioned problems was especially difficult in AndTweet because classes and methods had a large number of dependencies. As the number of dependencies increase, so does the likelihood that a method that the developer would like to offload depends on another method or class which contains a trouble spot, which, in turn prevents the offloading. It is possible to partly overcome this problem by creating smaller classes having fewer dependencies and internal state.
Our experiences clearly suggest that manually identifying the methods having offloading problems is non-trivial. It may include many cycles of trial-and-error and the only way to know whether all issues have been resolved is to test the application to see whether it works correctly or not. We think that it is essential to have tools that can automate this procedure and guide the programmers into developing more offloadable code.
\subsection{Energy consumption of local vs. offloaded Twitter}
\label{sec:andtweet_energy}
After eventually offloading Twitter communication, we measured the energy
consumption of non-modified and offloaded AndTweet. We tested offloading under two different network
conditions. In first setup the offloading server is in the same Wi-Fi
network as the phone. In the second scenario, the phone used 3G as the access network.
The main energy
consumers in mobile phones are CPU, I/O, display, and network interfaces.
We used the Monsoon Power
Monitor (www.msoon.com)
to measure the energy consumption of the phone. We also
collected the packet traces and used models to estimate the
energy consumed by the network interfaces. We used the model presented in \cite{yu10eenergy} to
estimate Wi-Fi energy consumption. As for 3G, we use a deterministic
power model which takes into account the different operating modes and
state transition controlled by inactivity timers according to the 3G
Radio Resource Control protocol\cite{balasubramanian09imc}. Power draw
of each state and the inactivity timer values were measured
beforehand. Based on the timer values and the observed traffic
patterns, the model deduces the time spent in each 3G radio state and
computes the energy consumption estimates. In the measurements and
the estimates, we excluded the one-time cost of transferring the
application image, because the applications could be, for example,
pre-installed to the offloading infrastructure.
\begin{table}[t]
\begin{scriptsize}
\begin{tabular}{ l l l }
Measurements & Wi-Fi(avg/stdev) & 3G(avg/stdev) \\
\hline
Total energy & 2.67/0.59 J & 3.92/1.97 J \\
Total en., offloaded & 25\% less/0.3 J & 18\% more/2.1 J \\
Network en., offloaded & 46\% less/0.04 J & 33\% more/2.0 J \\
Execution time & 2.69/0.59 s & 3.86/2.02 s\\
Exec. time, offloaded & 6\% more/0.5 s & 33\% more/2.1 s\\
Traffic & 7.7/0.8 kB & 6.0/2.1 kB \\
Traffic, offloaded & 17\% less/0.5 kB & 31\% less/1.8 kB \\
\end{tabular}
\caption{AndTweet offloading measurements for single refresh event. Total is measured and network is estimated energy consumption.}
\label{tab:andtweet-reload}
\end{scriptsize}
\end{table}
The results in Table \ref{tab:andtweet-reload} show energy consumption
of a single Twitter event which checks and fetches new tweets, and
then displays them. We
observe that offloading saves one fourth of the total energy consumed in the
Wi-Fi setup. As expected, the savings clearly come from having less
network traffic. However, the execution takes slightly longer
when offloaded, which increases the energy consumed by the
display. The results are very different when switching to 3G. More energy is consumed with the offloaded version even if less data
is transmitted and received. Because the RTT in 3G is an order of magnitude longer than with Wi-Fi, the remote invocation takes more time. Combined with long 3G inactivity timer values, this causes the network interface to be in
active high-power state (DCH) during the whole invocation. Execution time also has a big variance, because 3G latencies vary depending on network conditions and the current state of the radio when starting the transmission. Obviously, both network conditions and traffic patterns play a major role in the profitability of
offloading.
\section{SmartDiet}
\label{sec:toolkit}
\subsection{Overview}
In view of the experiences described above, we argue that the software
developer must be included in the offloading process. Furthermore,
a set of analysis tools are necessary to make this process
feasible. We therefore propose such a toolkit which we envision to comprise three tools: energy
analysis, constraint identification and software structure
analysis. We have designed and implemented prototypes of the
first two, but currently have only a vision of the third one.
The energy analysis tool finds and visualizes the parts of a
given application that could yield energy savings when offloaded. The
constraint identification tool automatically identifies constraints
in the source code, determines which methods can be offloaded as
such and points out trouble spots in the code. The idea for the
software structure analysis tool is that it analyzes the program code
to find opportunities to save energy through restructuring, for example by merging methods or classes.
The developer would first use the energy analysis tool to identify
the candidates for offloading. Next, the constraint identification
tool would be used to find and resolve problems in those candidates. Finally,
a structure analysis tool could be used to explore ways of making further
energy savings. In case such opportunities were found, the first two
tools would be applied again. This process will result in an application
that can be offloaded with larger energy savings by employing existing
frameworks.
\subsection{Energy analysis tool}
\label{sec:energy-usage-analysis-tool}
\subsubsection{Collecting measurements}
The energy analysis tool implements a measuring and modeling
setup for profiling and visualizing the energy consumption of a
given application. Since we collect data dynamically while the
program is running, the results reflect the specific usage scenario that is run.
The tool needs to collect three kinds of information:
the amount of computation, the amount of communication, and a trace of the
program execution flow to later produce class and method level
statistics for the developer.
Our tool collects information about the communication by capturing all IP traffic in the device, annotated with timestamps. To implement this, we
inject a kernel module using netfilter hooks
(www.netfilter.org) to access this data in
real-time. To collect CPU usage statistics, the tool uses
oprofiler (http://oprofile.sourceforge.net) in order
to do similar HPC-based computational energy profiling as in
\cite{yu10hpc} or analyze the /proc
filesystem as in
\cite{zhang10powertutor}. Both oprofiler and /proc filesystem
polling are in their unmodified form unable to gather data in a
sufficiently fine-grained fashion without adding a major overhead to the
overall system performance. In order to improve its performance, we plan to customize oprofiler to collect only the data we need. To track the program execution, the tool uses execution tracking
features in Android Debug Monitor Server
(DDMS)
which produces a trace of all classes and methods executed during the
run. In addition, we implemented minor modifications to the Dalvik
virtual machine to annotate the traces with system-wide timestamps, which
can be matched to those of our other measurements.
\begin{figure}[t]
\centering
\includegraphics[width=240pt, height=150pt]{methods-and-packets-annotated.pdf}
\caption{TCP packets and network-related method calls for test application fetching an HTML page.}
\label{fig:methods-and-packets}
\end{figure}
\begin{figure*}[t]
\centering
\includegraphics[width=500pt]{simpleapp-fixed-annotations.pdf}
\caption{Network usage graph for test application, which fetches an HTML page. Numbers show how many times methods have been invoked during the whole procedure and how much energy it consumed.}
\label{fig:simpleapp}
\end{figure*}
\subsubsection{Matching packets to method trace}
SmartDiet associates each packet in the collected packet trace
to an individual method in the program execution trace.
The tool starts by dividing the execution trace into threads
and packet trace into separate flows (TCP connection or UDP
flow). Furthermore, only network-related method calls are
filtered for each thread. We now have two separate time series:
network-related method calls of each thread and packet arrival events of each
flow (see Figure \ref{fig:methods-and-packets} for
illustration). These series are compared by computing
cross-correlations in order to associate each flow to a particular
thread which is generating that traffic. The idea is that each
network-related method is associated with the corresponding
packets. Finally, each packet of a flow is associated to the closest
(in time) method call of the corresponding thread. This way, the
tool generates a method trace of the program execution annotated with
information about the methods that caused network traffic.
Figure \ref{fig:methods-and-packets} shows part of the traces of a
simple Android test application that performs an HTTP GET request when
a button is clicked. The thread executing the HTTP request correlates
strongly with the packet trace. Another thread in the figure has a single
network-related call. It is the garbage collector thread
running finalization for a network-related object that is no longer
used. Since it correlates weakly with the packet trace, no packets are
associated with it.
\subsubsection{Visualizing the network usage}
Program execution for each thread can be viewed as a hierarchical call
tree, where a method calls another method which calls another and so
on. Our tool reconstructs this tree, carrying along the information of
the detected network usage. It then aggregates the traffic of the
nodes up in the tree, so that the root method, where the execution
starts, gets associated with all packets that have been sent or
received within each thread.
Figure \ref{fig:simpleapp} is an example graph, automatically produced by SmartDiet, for a simple test case requesting an HTML document over HTTP. Traffic is cumulatively assigned to the MainActivity.onClick
method and from there on, divided between various library functions
that open the connection, send an HTTP request, receive the response
and finally close the connection. At each step, we present the number
of calls made and traffic statistics alongside with
energy consumption estimates based on the models described in
Section \ref{sec:andtweet_energy}.
The energy usage estimate is shown as a range between two values,
because in estimation one has to make assumptions about the
dependence of traffic in a single method to the traffic of the rest
of the program. Let $N$ be the full packet trace,
$N_{method}$ the part of the packet trace associated with a given
method and $N_{rest}$ everything from $N$ that does not belong to
$N_{method}$. $E(n)$ is the energy-estimate for a packet trace $n$ and hence $E(N)$ energy-estimate for the whole program.
Assuming that all other traffic is left unmodified as we offload the
traffic of a single method, $E(N_{rest})$ would be the remaining
energy consumption. The energy savings would hence be $E(N) -
E(N_{rest})$. If we assume that all other traffic is independent of
our method, we can calculate the energy saved as $E(N_{method})$. The latter estimate is always larger than
the former, because sending multiple packets together is always
energy-wise cheaper than sending them separately. Because we do not estimate the dependencies between traffic caused by different methods, we take the
former as the lower bound of energy savings $E_{min}$ and the latter
as the upper bound $E_{max}$, and let the developer do further reasoning.
When we tried this approach on more complex
applications, mainly AndTweet and
ConnectBot\footnote{\url{http://code.google.com/p/connectbot/}}, we identified
a caveat: Using the method trace information only, we are not always able to accurately track the data flow between threads. If an application uses
a threading network library--the case especially with
ConnectBot--our aggregated results will show energy consumption only within
the library code. The developer, however, is mostly interested in
those parts of the application code that actually generated this
traffic. Section \ref{sec:challenges} discusses potential solutions to this problem of execution tracking.
\subsection{Constraint identification tool}
\label{sec:constraint-identification-tool}
SmartDiet's constraint identification performs the analysis on the application source code. For each method in the application, it points out problems that can prevent offloading of that method unless the code is modified. SmartDiet currently focuses on heuristics that identify problems associated with our offloading setup, which is using the Android platform and Java Serialization API to implement the remote execution of methods. Nevertheless, similar heuristics can be crafted to other remote execution mechanisms, including the Android Parcelable mechanism used in Cuckoo \cite{kemp10cuckoo}, .NET serialization used in MAUI\cite{cuervo10maui}, or even the virtual machine based thread migration used in CloneCloud \cite{chun11clonecloud}, because they all set some restrictions on what kind of methods can be offloaded.
\subsubsection{Identifying hardware constraints: access to local resources}
The first set of constrained methods are those that require access to the hardware of the local device. If a method accesses one of the constrained Android system APIs, it cannot be offloaded unless the code structure is changed. We currently identify method as having this constraint if it tries to show, for instance, notifications to the user, update anything on the screen, vibrate the phone, access the Bluetooth, wifi or usb subsystem, and so on. We have identified a total set of 20 constrained subsystems.
\subsubsection{Identifying software constraints: unexpected behavior or limitations in migratability}
Our second set of constrained methods are those that cannot be migrated to the remote server at all due to migration mechanism requirements, or those that cause unexpected behavior when executed remotely due to inconsistent states between local and remote execution environments.
Migration limitations are specific to the mechanism used, which in our case is the Java serialization APIs. For a method to be migratable, its encapsulating class as well as arguments and return type must implement the Java serializable interface. SmartDiet finds all methods that adhere to this criterion to show which ones can be directly migrated.
We find that only a fraction of analyzed methods implement the serializable requirements. We therefore also calculate the number of methods which could be modified to be serializable with a few minor changes. In this category, we include all methods whose encapsulating class, as well as arguments and return types, are convertible to serializable with the following criteria: A type is considered convertible to serializable, if all of its supertypes and member classes are either directly serializable or belong to this application's codebase and can be also converted to serializable using this same principle recursively. We exclude the library code, because for instance the Android SDK code cannot usually be easily modified, even if the changes would be simple.
Regarding unexpected behavior, SmartDiet finds all methods that access the local file system using either Android's SharedPreferences mechanism or Java's File class. ThinkAir does not synchronize the file system, thus files in the remote server are not the same as those in the device, which will often cause unexpected behavior for the program. This principle can be extended to find problems related to other non-synchronized resources as well.
A potential solution to the synchronization issues is a system which automatically synchronizes the relevant state. This is notably a hard problem on its own, but in the context of offloading we are also concerned about energy usage of the synchronization. Until efficient automatic solutions are presented, the developer must be assisted in overcoming the problems manually.
Based on our experiences with AndTweet, we believe that an even richer
set of rules could be established to detect and classify different
types of remote execution issues, as well as provide suggestions to the
developer for overcoming them. For example, if a class contains
an instance of the non-serializable HttpClient class, as described in
Section \ref{sec:constraints-of-remote-execution}, we might suggest
removing the member instance and replacing it with a new instance of
HttpClient created on-the-fly every time it is needed.
\subsubsection{Statistics from open source programs}
\label{sec:program-statistics}
\begin{table}
\begin{scriptsize}
\begin{tabular}{ p{4cm} l l l l }
Statistic & Median & Min & Max \\
\hline
Number of methods & 431 & 121 & 4411 \\
Directly migratable & 0.17\% & 0.00\% & 3.70\% \\
Migratable with minor changes & 15.7\% & 0.00\% & 46.8\% \\
Hardware access constraints & 14.2\% & 2.28\% & 41.3\% \\
Potential unexpected behavior & 10.7\% & 0.00\% & 30.3\% \\
because of access to file system & & & \\
\end{tabular}
\end{scriptsize}
\caption{Constraint statistics for 16 open source applications.}
\label{tab:heuristics}
\end{table}
Our tool analyzes source code, which restricts us to analyzing
open-source software. Unfortunately the most popular applications on
Android Market are closed-source. To find similar applications, we
went through numerous Android application listings, and selected any programs that are
non-trivial in size and include either communication or non-trivial
computation\footnote{We used Android Market, popular source code sharing site Github, Wikipedia and numerous other sources}. Additionally we analyze also some platform applications
that are shipped with Android operating system, like the web browser.
We ran our constraint analysis tool on this set of programs. Results are shown in Table \ref{tab:heuristics}. Maximum of 3\% of methods are directly migratable, but SmartDiet can point out changes to source code which will enable the migration of 15 to 47\% of methods for offloading using ThinkAir. SmartDiet can also guide the developer into fixing the issues regarding hardware or filesystem access.
However, even after these changes, an overwhelming majority of methods have migration issues. Here's where the third tool in our toolset, one that performs structural analysis, would be helpful. We could guide the developer in making larger structural changes, which would enable new portions of the application to be offloaded, in an energy-efficient fashion.
\section{Remaining challenges}
\label{sec:challenges}
In our current prototype we have implemented dynamic measurement of communication energy usage. Existing solutions can be applied to estimate computation and display energy consumption\cite{yu10hpc,zhang10powertutor} given just that time stamping and high enough sampling rates can be supported. We plan to integrate CPU usage measurement and estimation to the same toolkit. We can find CPU hot spots in a similar fashion as we have done for network transmission.
As discussed in Section \ref{sec:energy-usage-analysis-tool}, we encountered problems when analyzing the execution flow of complex programs, relying heavily on threading. We are currently looking into the TaintDroid \cite{enck10taintdroid} system, which modifies the Android Java virtual machine in such a way that data flows can be tracked, and would like to extend our toolkit using similar mechanisms for data and execution flow tracking over thread boundaries.
Regarding our network energy usage modeling, we currently use only the traffic as input. In order to explicitly show the impact of network conditions, the power models can be improved by using metrics reflecting network conditions as parameters.
Conserning the application structure analysis tool, we believe it would be beneficial to investigate which kind of programming styles and application structures best suit offloading. Combining the results of our heuristics to standard object-oriented code quality measures, like coupling and cohesion, could yield interesting results and insights into how we can enhance the effectiveness of offloading. This can also lead to suggestions how to improve programs regarding offloadability.
\section{Conclusions}
In this paper we studied feasibility and potential energy savings that are achievable utilizing method-level offloading, especially in network-intensive mobile applications. We used an open source Twitter client as an example to show that completely automated offloading often misses opportunities for saving energy and may also lead to various failures during execution. For this reason, we argue that the developer should participate actively in the offloading process. To this end, we propose an offloading analysis toolkit called SmartDiet. It analyzes Android application source code and collects run-time information in order to help developers in identifying potential energy savings and trouble spots in their program code so that existing offloading frameworks can be more efficiently utilized. SmartDiet has shown promising results in helping the development of offloadable code, although some challenges still remain to complete all the features we believe would be useful.
\bibliographystyle{abbrv}
|
\section{Introduction}
In 1971, Viterbi \cite{Viterbi1971} published a nowadays classical upper bound on the bit error probability $P_\text{b}$ for Viterbi decoding, when convolutional codes are used to communicate over the binary symmetric channel (BSC). This bound was derived from the extended path weight enumerators, obtained using a signal flow chart technique for convolutional encoders. Later, van de Meeberg \cite{Meeberg1974} used a very clever observation to tighten Viterbi's bound for large signal-to-noise ratios (SNRs).
The challenging problem of deriving an expression for the \textit{exact} (decoding) bit error probability was first addressed by Morrissey in 1970 \cite{Morrissey1970} for a suboptimal feedback decoding algorithm. He obtained the same expression for the exact bit error probability for the rate $R=1/2$, memory $m=1$ ($2$-state) convolutional encoder with generator matrix $G(D) = (1 \quad 1+D)$ that Best et al. \cite{Best1995} obtained for Viterbi decoding. Their method is based on considering a Markov chain of the so-called metric states of the Viterbi decoder; an approach due to Burnashev and Cohn \cite{Burnashev1990}. An extension of this method to the rate $R=1/2$ memory $m=2$ ($4$-state) convolutional encoder with generator matrix $G(D) = (1+D^2 \quad 1+D+D^2)$ was published by Lentmaier et al. \cite{Lentmaier2004}.
In this paper we use a different and more general approach to derive a closed form expression for the exact (decoding) bit error probability for Viterbi decoding of convolutional encoders, when communicating over the BSC as well as the quantized additive white Gaussian noise (AWGN) channel. Our new method allows the calculation of the exact bit error probability for more complex encoders in a wider range of code rates than the methods of \cite{Best1995} and \cite{Lentmaier2004}. By considering a random tie-breaking strategy, we average the information weights over the channel noise sequence and the sequence of random decisions based on coin-flippings (where the coin may have more than two sides depending on the code rate). Unlike the backward recursion in \cite{Best1995} and \cite{Lentmaier2004}, the bit error probability averaged over time is obtained by deriving and solving a recurrent matrix equation for the average information weights at the current and previous states of a trellis section when the maximum-likelihood branches are decided by the Viterbi decoder at the current step.
To illustrate our method, we use a rate $R=2/3$ systematic convolutional $2$-state encoder whose minimal realization is given in observer canonical form, since this encoder is both general and simple.
In \secref{recurrent_equation}, the problem of computing the exact bit error probability is reformulated via the average information weights. A recurrent matrix equation for these average information weights is derived in \secref{inf_weights} and solved in \secref{solving_the_recurrent_equation}. In \secref{examples}, we give additional examples of rate $R=1/2$ and $R=2/3$ encoders of various memories. Furthermore, we analyze a rate $R=1/2$ $4$-state encoder used to communicate over the quantized additive white Gaussian noise (AWGN) channel and show an interesting result that would be difficult to obtain without being able to calculate the exact bit error probability.
Before proceeding, we would like to emphasize that the bit error probability is an encoder property, neither a generator matrix property nor a convolutional code property.
\section{Problem Formulation via the Average Information Weights}\label{sec:recurrent_equation}
Assume that the all-zero sequence is transmitted over a BSC with crossover probability $p$ and let $W_t(\sigma)$ denote the weight of the information sequence corresponding to the code sequence decided by the Viterbi decoder at state $\sigma$ and time instant $t$. If the initial values $W_0(\sigma)$ are known, then the random process $W_t(\sigma)$, $t=0,1,2,\ldots$, is a function of the random sequence of the received $c$-tuples $\bs{r}_\tau$, $\tau=0,1,\ldots,t-1$, and the coin-flippings used to resolve ties.
Our goal is to determine the mathematical expectation of the random variable $W_t(\sigma)$ over this ensemble, since for rate $R=b/c$ minimal convolutional encoders the bit error probability can be computed as the limit
\begin{IEEEeqnarray}{rCl}
\label{eq: viterbi_bit_error_probability}
P_{\text{b}} & = & \lim_{t \to \infty} \frac{E\left[ W_t(\sigma = 0)\right]}{tb}
\end{IEEEeqnarray}
assuming that this limit exists.
\textit{Remark.} If we consider nonminimal encoders, all states equivalent to the all-zero state have to be also taken into account.
We consider encoder realizations in both controller and observer canonical form and denote the encoder states by $\sigma$, $\sigma \in \{ 0, 1, \ldots, \abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}-1 \}$, where {\footnotesize\raisebox{0.1em}{$\sum$}} is the set of all possible encoder states.
During the decoding step at time instant $t+1$ the Viterbi algorithm computes the cumulative Viterbi branch metric vector $\bs{\mu}_{t+1} = \left( \mu_{t+1}(0) \enspace \mu_{t+1}(1) \ldots \mu_{t+1}(\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}-1) \right)$ for the time instant $t+1$ using the vector $\bs{\mu}_t$ and the received $c$-tuple $\bs{r}_t$. It is convenient to normalize the metrics such that the cumulative metrics at every all-zero state will be zero, that is, we subtract the value $\mu_t(0)$ from $\mu_t(1), \mu_t(2), \ldots, \mu_t(\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}-1)$ and introduce the normalized cumulative branch metric vector
\begin{IEEEeqnarray}{rCl}
\bs{\phi}_t & = & \Big( \phi_t(1) \enspace \phi_t(2) \ldots \phi_t(\abs{\text{\tiny\raisebox{0.25em}{$\sum$}}}-1) \Big) \nonumber \\
& = & \Big( \mu_t(1)\!-\!\mu_t(0) \enspace \mu_t(2)\!-\!\mu_t(0) \ldots \mu_t(\abs{\text{\tiny\raisebox{0.25em}{$\sum$}}}-1)\!-\!\mu_t(0) \Big) \nonumber
\end{IEEEeqnarray}
For example, for a $2$-state encoder we obtain the scalar
\begin{IEEEeqnarray}{rCl}
\phi_t & = & \phi_t(1) \nonumber
\end{IEEEeqnarray}
while for a $4$-state encoder we have the vector
\begin{IEEEeqnarray}{rCl}
\bs{\phi}_t & = & \Big( \phi_t(1) \enspace \phi_t(2) \enspace \phi_t(3) \Big) \nonumber
\end{IEEEeqnarray}
The elements of the random vector $\bs{\phi}_t$ belong to a set whose cardinality $M$ depends on the channel model, encoder structure, and the tie-breaking rule. Enumerating the vectors $\bs{\phi}_t$ by numbers $\phi_{t}$ which are random variables taking on $M$ different integer values $\phi^{(0)},\,\phi^{(1)},\ldots,\,\phi^{(M-1)}$, the sequence of numbers $\phi_{t}$ forms an $M$-state Markov chain $\Phi_{t}$ with transition probability matrix $\Phi = \left( \phi_{jk} \right)$, where
\begin{equation}
\phi_{jk} = \text{Pr} \left( \phi_{t+1} = \phi^{(k)} \;\middle\vert\; \phi_t = \phi^{(j)} \right)
\end{equation}
\begin{figure}[t]
\centering
\includegraphics{ocf_realization-external}
\caption{\label{fig:ocf_realization}A minimal encoder for the generator matrix given in equation \eqref{eq:ocf_encoding_matrix_R23}.}
\end{figure}
\begin{figure*}[t]
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_a-external.pdf}\hspace{-14.2272pt}}
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_b-external.pdf}\hspace{-14.2272pt}}
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_c-external.pdf}\hspace{-14.2272pt}}
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_d-external.pdf}\hspace{-14.2272pt}} \\[-2mm]
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_e-external.pdf}\hspace{-14.2272pt}}
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_f-external.pdf}\hspace{-14.2272pt}}
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_g-external.pdf}\hspace{-14.2272pt}}
\subfloat[]{\hspace{-15.9339pt}\includegraphics{ocftrellis_h-external.pdf}\hspace{-14.2272pt}}
\caption{\label{fig:eight_trellis_sections}Eight (of a total of $40$) trellis sections for the rate $R=2/3$, $2$-state encoder in \figref{ocf_realization}.}
\end{figure*}
Let $\bs{W}_t$ be the vector of information weights at time instant $t$ that depends both on the $\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ encoder states $\sigma_t$ and on the $M$ normalized cumulative metrics $\phi_t$; that is, $\bs{W}_t$ is expressed as the following vector with $M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ entries
\begin{IEEEeqnarray}{rclccccr}
\bs{W}_t & = & \Big( & \bs{W}_t(\sigma=0) \enspace & \bs{W}_t(\sigma=1) & \enspace \ldots \enspace & \bs{W}_t(\sigma=\abs{\text{\tiny\raisebox{0.25em}{$\sum$}}}\!-\!1) \Big) \IEEEeqnarraynumspace
\end{IEEEeqnarray}
where
\begin{IEEEeqnarray}{rclcccr}
\bs{W}_t(\sigma) & = & \Big( & W_t(\phi^{(0)}, \sigma) \enspace & W_t(\phi^{(1)}, \sigma) & \enspace \ldots \enspace & W_t(\phi^{(M-1)}, \sigma) \Big) \IEEEeqnarraynumspace
\end{IEEEeqnarray}
Then \eqref{eq: viterbi_bit_error_probability} can be rewritten as
\begin{IEEEeqnarray}{rCl}
P_{\text{b}} & = & \lim_{t \to \infty} \frac{E[W_t(\sigma = 0)]}{tb} = \lim_{t \to \infty} \frac{\sum_{i=0}^{M\!-\!1} E[W_t(\phi^{(i)}, \sigma = 0)]}{tb} \nonumber \\[0.5mm]
& = & \lim_{t \to \infty} \frac{E[\bs{W}_t(\sigma = 0)] \bs{1}^\text{T}_{1,M}}{tb} = \lim_{t \to \infty} \frac{\bs{w}_t(\sigma = 0) \bs{1}^\text{T}_{1,M}}{tb}\nonumber \\[0.5mm]
& = & \lim_{t \to \infty} \frac{\bs{w}_t}{tb} \big( \bs{1}_{1,M}\, \bs{0}_{1,M}\ldots\,\bs{0}_{1,M})^{\text{T}} \label{eq: viterbi_bit_error_probability_w}
\end{IEEEeqnarray}
where $\bs{1}_{1,M}$ and $\bs{0}_{1,M}$ denote the all-one and the all-zero row vectors of length $M$, respectively, $\bs{w}_{t}$ represents the length $M \abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ vector of the average information weights, while the length $M$ vector of average information weights at the state $\sigma$ is given by $\bs{w}_t(\sigma)$. Note that the mathematical expectations in \eqref{eq: viterbi_bit_error_probability_w} are computed over the sequences of channel noises and coin-flipping decisions.
To illustrate the introduced notations, we use the rate $R=2/3$ memory $m=1$, overall constraint length $\nu=2$, minimal encoder with systematic generator matrix
\begin{equation}
G(D) =
\left(\begin{array}{ccc}
1 & 0 & 1+D \\
0 & 1 & 1+D
\end{array}\right) \label{eq:ocf_encoding_matrix_R23}
\end{equation}
It has a $2$-state realization in observer canonical form as shown in \figref{ocf_realization}.
\begin{figure}[t]
\centering
\includegraphics{markov_chain-external}
\caption{\label{fig:state_transition_diagram}Illustration of the $5$-state Markov chain formed by the sequences of normalized cumulative metric states $\phi_t$.}
\vspace{-1em}
\end{figure}
Assuming that the normalized cumulative metric state is $\phi_t = 0$, we obtain the eight trellis sections given in \figref{eight_trellis_sections}. These trellis sections yield the normalized cumulative metric states $\left\{ -1, 0, 1 \right\}$. Using $\phi_t = -1$ and $\phi_t = 1$, we obtain $16$ additional trellis sections and two additional normalized cumulative metric states $\left\{ -2, 2 \right\}$. From the metrics $\phi_t = -2$ and $\phi_t = 2$, we get another $16$ trellis sections but those will not yield any new metrics. Thus, in total we have $M=5$ normalized cumulative metric states $\phi_t \in \left\{ -2, -1, 0, 1, 2\right\}$. Together with the eight different received triples, $\bs{r}_t = 000$, $001$, $010$, $100$, $011$, $101$, $110$, $111$, they correspond to in total $40$ different trellis sections. The bold branches in \figref{eight_trellis_sections} correspond to the branches decided by the Viterbi decoder at time instant $t+1$. When we have more than one branch with maximum normalized cumulative metric entering the same state, we have a tie which we, in our analysis, resolve by fair coin-flipping.
Hence, the normalized cumulative metric $\Phi_t$ is a $5$-state Markov chain with transition probability matrix $\Phi = \left( \phi_{jk} \right)$, $1 \leq j,k \leq 5$.
From the four trellis sections, (a), (b), (g), and (h), in \figref{eight_trellis_sections} we obtain
\begin{IEEEeqnarray}{rCl}
\phi_{0(-1)} & = & \text{Pr}\left( \bs{r}_t = 000 \right) + \text{Pr}\left( \bs{r}_t = 001 \right) \nonumber \\
& & \,+\,\text{Pr}\left( \bs{r}_t = 110 \right) + \text{Pr}\left( \bs{r}_t = 111 \right) \nonumber \\
& = & q^3 + pq^2 + p^2 q + p^3 = p^2 + q^2
\end{IEEEeqnarray}
while the four trellis sections, (c), (d), (e), and (f), yield
\begin{equation}
\phi_{01} = pq^2 + pq^2 + p^2q + p^2q = 2pq
\end{equation}
where $q=1-p$.
Similarly, we can obtain the remaining transition probabilities from the $32$ trellis sections not included in \figref{eight_trellis_sections}. Their transition probability matrix follows as
\vspace{1em}
\begin{IEEEeqnarray}{c}
\Phi = \fharray{2em}{2em}{1em}{1em}{1em}{l@{\hspace{0em}}ccccc@{\hspace{0em}}c}{
& \sm{-2} & \sm{-1} & \sm{0} & \sm{1} & \sm{2} & \phi^{(k)}\\
\sm{-2} & q^3 + p^2q & 0 & p^3 + 3pq^2 & 0 & 2p^2q \\
\sm{-1} & q^3 + p^2q & 0 & p^3 + 3pq^2 & 0 & 2p^2q \\
\sm{\phantom{-}0} & 0 & p^2 + q^2 & 0 & 2pq & 0 \\
\sm{\phantom{-}1} & p^3 + pq^2 & 0 & q^3 + 3p^2q & 0 & 2p^2q \\
\sm{\phantom{-}2} & p^3 + pq^2 & 0 & q^3 + 3p^2q & 0 & 2p^2q \\
\phi^{(j)}
}\vspace{1em}\IEEEeqnarraynumspace
\end{IEEEeqnarray}
whose metric state Markov chain is shown in \figref{state_transition_diagram}.
Let $\bs{p}_t$ denote the probabilities of the $M$ different normalized cumulative metric values of $\Phi_t$, that is, $\phi_t \in \{ \phi^{(0)}, \phi^{(1)}, \ldots, \phi^{(M-1)} \}$. Their stationary distribution is denoted $\bs{p}_{\infty} = ( p_{\infty}^{(0)} \enspace p_{\infty}^{(1)} \ldots p_{\infty}^{(M-1)} )$ and is determined as the solution of, for example, the first $M-1$ equations of
\begin{IEEEeqnarray}{rCl}
\label{steady-state}
\bs{p}_{\infty} \Phi & = & \bs{p}_{\infty}
\end{IEEEeqnarray}
and
\vspace{-1em}
\begin{IEEEeqnarray}{rCl}
\sum_{i=0}^{M-1} p_{\infty}^{(i)} & = & 1
\end{IEEEeqnarray}
For the $2$-state convolutional encoder with generator matrix \eqref{eq:ocf_encoding_matrix_R23} we obtain
{\setlength\arraycolsep{0.1em}
\renewcommand{\arraystretch}{1.2}
\small
\begin{IEEEeqnarray*}{c}
\bs{p}_{\infty}^\text{T} = \frac{1}{1 -p +10p^2 -20p^3 +20p^4 -8p^5} \\
\times \left(\begin{array}{ccccccccccccccc}
1 & + & 7p & - & 28p^2 & + & 66p^3 & - & 100p^4 & + & 96p^5 & - & 56p^6 & + & 16p^7 \\
& - & 3p & + & 16p^2 & - & 46p^3 & + & 80p^4 & - & 88p^5 & + & 56p^6 & - & 16p^7 \\
& - & 3p & + & 10p^2 & - & 20p^3 & + & 20p^4 & - & 8p^5 \\
& & & - & 6p^2 & + & 26p^3 & - & 60p^4 & + & 80p^5 & - & 56p^6 & - & 16p^7 \\
& & & - & 2p^2 & - & 6p^3 & + & 40p^4 & - & 72p^5 & + & 56p^6 & - & 16p^7
\end{array}\right)
\end{IEEEeqnarray*}}%
In order to compute the exact bit error probability according to \eqref{eq: viterbi_bit_error_probability_w}, it is necessary to determine $\bs{w}_{t}(\sigma=0)$. In the next section we will derive a recurrent matrix equation for the average information weights and illustrate how to obtain its components using as an example the rate $R=2/3$ memory $m=1$ minimal encoder determined by \eqref{eq:ocf_encoding_matrix_R23}.
\section{Computing the Vector of Average Information Weights}\label{sec:inf_weights}
The vector $\bs{w}_t$ describes the dynamics of the information weights when we proceed along the trellis and satisfies the recurrent equation
\begin{IEEEeqnarray}{C}
\label{eq: recurrent_equation}
\left\{ \,
\begin{IEEEeqnarraybox}[\IEEEeqnarraystrutmode\IEEEeqnarraystrutsizeadd{2pt}{2pt}][c]{rCl}
\bs{w}_{t+1} & = & \bs{w}_t A + \bs{b}_t B \\
\bs{b}_{t+1} & = & \bs{b}_t \Pi
\end{IEEEeqnarraybox} \right.
\end{IEEEeqnarray}
where $A$ and $B$ are $M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}} \times M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ nonnegative matrices, $\Pi$ is an $M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}} \times M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ stochastic matrix, and $\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}=2^m$. Both matrices consist of $\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}} \times \abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ submatrices $A_{ij}$ and $B_{ij}$ of size $M \times M$, respectively, where the former satisfy
\begin{equation}
\sum_{i=0}^{\abs{\text{\tiny\raisebox{0.25em}{$\sum$}}}-1} A_{ij} = {\Phi-1}, \enspace j=0,1,\ldots,\abs{\text{\tiny\raisebox{0.25em}{$\sum$}}}-1 \label{eq:row_sum}
\end{equation}
since we consider only encoders for which every encoder state is reachable with probability $1$.
The matrix A represents the linear part of the affine transformation of the information weights while the matrix $B$ describes their increments. The submatrices $A_{ij}$ and $B_{ij}$ describe the updating of the average information weights if the transition from state $i$ to state $j$ exists; and are zero otherwise. Moreover, the vector $\bs{b}_t$ of length $M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ is the concatenation of $\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ stochastic vectors $\bs{p}_t$, and hence the $M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}} \times M\abs{\text{\footnotesize\raisebox{0.1em}{$\sum$}}}$ matrix~$\Pi$ follows as
\begin{IEEEeqnarray}{rCl}
\Pi = \left(\begin{array}{cccc}
\Phi & 0 & \ldots & 0 \\
0 & \Phi & \ldots & 0 \\
\vdots & \vdots & \ddots & \vdots \\
0 & 0 & \ldots & \Phi
\end{array}\right)
\end{IEEEeqnarray}
For simplicity, we choose the initial value of the vector of the information weights to be
\begin{IEEEeqnarray}{rCl}
\label{eq: initial_value_w}
\bs{w}_0 & = & \bs{0}
\end{IEEEeqnarray}
Continuing the previous example, we will illustrate how the $10 \times 10$ matrices $A$ and $B$ can be obtained directly from all $40$ trellis sections. For example, the eight trellis sections in \figref{eight_trellis_sections} determine all transitions from $\phi_t = 0$ to either $\phi_{t+1} = -1$ or $\phi_{t+1} = 1$.
To be more specific, consider all transitions from $\sigma_t = 0$ and $\phi_t = 0$ to $\sigma_{t+1} = 0$ and $\phi_{t+1} = -1$, as shown in \figref{eight_trellis_sections}(a), (b), (g), and (h). Only \figref{eight_trellis_sections}(a) and (g) have transitions decided by the Viterbi algorithm, which are $\bs{v}_t = 000$ in \figref{eight_trellis_sections}(a) and $\bs{v}_t = 110$ in \figref{eight_trellis_sections}(g), and thus the entry $\sigma_t = 0$, $\phi_t = 0$, $\sigma_{t+1} = 0$, $\phi_{t+1} = -1$ in matrix $A$ follows as
\begin{equation}
\text{Pr}\left( \bs{r}_t = 000 \right) + \text{Pr}\left( \bs{r}_t = 110 \right) = q^3 + p^2q \nonumber
\end{equation}
and in matrix $B$ as
\begin{IEEEeqnarray}{rCl}
\IEEEeqnarraymulticol{3}{l}{\beta\left(000\right)\text{Pr}\left( \bs{r}_t = 000 \right) + \beta\left(110\right)\text{Pr}\left( \bs{r}_t = 110 \right)} \nonumber \\
\quad & = & 0 + 2p^2q = 2p^2q \nonumber
\end{IEEEeqnarray}
where $\beta\left( \bs{v}_t \right)$ denotes the number of information $1$s corresponding to $\bs{v}_t$. Since we use coin-flipping to resolve ties, we obtain that the entry $\sigma_t = 0$, $\phi_t = 0$, $\sigma_{t+1} = 0$, $\phi_{t+1} = 1$ (\figref{eight_trellis_sections}(c) and (d)) in matrix $A$ is
\begin{IEEEeqnarray}{rCl}
\IEEEeqnarraymulticol{3}{l}{\frac{1}{2} \text{Pr} \left( \bs{r}_t = 010 \right) + \frac{1}{2} \text{Pr} \left( \bs{r}_t = 010 \right)} \nonumber \\
\IEEEeqnarraymulticol{3}{l}{\, + \frac{1}{2} \text{Pr} \left( \bs{r}_t = 100 \right) + \frac{1}{2} \text{Pr} \left( \bs{r}_t = 100 \right)} \nonumber \\
\quad & = & \frac{1}{2} pq^2 + \frac{1}{2} pq^2 + \frac{1}{2} pq^2 + \frac{1}{2} pq^2 = 2pq^2 \nonumber
\end{IEEEeqnarray}
and in matrix $B$
\begin{IEEEeqnarray}{rCl}
\IEEEeqnarraymulticol{3}{l}{\frac{1}{2} \beta\left(000\right)\text{Pr} \left( \bs{r}_t = 010 \right) + \frac{1}{2} \beta\left(110\right) \text{Pr} \left( \bs{r}_t = 010 \right)} \nonumber \\
\IEEEeqnarraymulticol{3}{l}{\, + \frac{1}{2} \beta\left(000\right) \text{Pr} \left( \bs{r}_t = 100 \right) + \frac{1}{2} \beta\left(110\right) \text{Pr} \left( \bs{r}_t = 100 \right) } \nonumber \\
\quad & = & \frac{1}{2} \cdot 0 + \frac{1}{2} \cdot 2 pq^2 + \frac{1}{2} \cdot 0 + \frac{1}{2} \cdot 2 pq^2 = 2pq^2 \nonumber
\end{IEEEeqnarray}
Similarly the entry $\sigma_t = 1$, $\phi_t = 0$, $\sigma_{t+1} = 0$, $\phi_{t+1} = -1$ (\figref{eight_trellis_sections}(b) and (h)) in matrix $A$ is
\begin{equation*}
pq^2 + p^3
\end{equation*}
and in matrix $B$
\begin{equation*}
0 + 2p^3 = 2p^3
\end{equation*}
Finally, the entry $\sigma_t = 1$, $\phi_t =0$, $\sigma_{t+1} = 0$, $\phi_{t+1} = 1$ (\figref{eight_trellis_sections}(e) and (f)) in matrix $A$ is given by
\begin{equation*}
\frac{1}{2} p^2 q + \frac{1}{2} p^2q + \frac{1}{2} p^2 q + \frac{1}{2} p^2 q = 2p^2q
\end{equation*}
and in matrix $B$ by
\begin{equation*}
\frac{1}{2} \cdot 0 + \frac{1}{2} \cdot 2 p^2 q + \frac{1}{2} \cdot 0 + \frac{1}{2} \cdot 2 p^2 q = 2p^2q
\end{equation*}
The trellis sections in \figref{eight_trellis_sections} determine also the entries for the transitions $\sigma_t = 0$, $\phi_t = 0$, $\sigma_{t+1} = 1$, $\phi_{t+1} = -1$ and $\sigma_t = 0$, $\phi_t = 0$, $\sigma_{t+1} = 1$, $\phi_{t+1} = 1$ as well as the transitions $\sigma_t =1$, $\phi_t = 0$, $\sigma_{t+1} = 1$, $\phi_{t+1} = -1$ and $\sigma_t = 1$, $\phi_t = 0$, $\sigma_{t+1} = 1$, $\phi_{t+1} = 1$.
The remaining transitions with $\phi_t = 0$ are never decided by the Viterbi algorithm, and hence the corresponding entries are zero. The eight trellis sections in \figref{eight_trellis_sections} yield $20$ entries in the matrices $A$ and $B$, while the $32$ trellis sections not shown in \figref{eight_trellis_sections} yield the remaining $80$ entries. For the convolutional encoder shown in \figref{ocf_realization} we obtain
\begin{equation}
A = \left( \begin{array}{cc} A_{00} & A_{01} \\ A_{10} & A_{11} \end{array} \right)
\end{equation}
where
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
A_{00} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & q^3\!+\!p^2q & 0 & p^3\!+\!3pq^2 & 0 & 2p^2q \\
\sm{-1} & q^3\!+\!p^2q & 0 & \frac{1}{2}p^3\!+\!\frac{5}{2}pq^2 & 0 & p^2q \\
\sm{0} & 0 & q^3\!+\!p^2q & 0 & 2pq^2 & 0 \\
\sm{1} & 0 & 0 & \frac{1}{2}q^3\!+\!\frac{1}{2}p^2q & 0 & pq^2 \\
\sm{2} & 0 & 0 & 0 & 0 & 0
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
A_{01} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & p^2q\!+\!q^3 & 0 & p^3\!+\!3pq^2 & 0 & 2p^2q \\
\sm{-1} & \frac{1}{2}p^3\!+\!\frac{1}{2}pq^2 & 0 & p^2q & 0 & 0 \\
\sm{0} & 0 & p^3\!+\!pq^2 & 0 & 2p^2q & 0 \\
\sm{1} & \frac{1}{2}q^3\!+\!\frac{1}{2}p^2q & 0 & p^3\!+\!2pq^2 & 0 & 2p^2q \\
\sm{2} & 0 & 0 & 0 & 0 & 0
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
A_{10} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & 0 & 0 & 0 & 0 & 0 \\
\sm{-1} & 0 & 0 & \frac{1}{2}p^3\!+\!\frac{1}{2}pq^2 & 0 & p^2q \\
\sm{0} & 0 & p^3\!+\!pq^2 & 0 & 2p^2q & 0 \\
\sm{1} & p^3\!+\!pq^2 & 0 & \frac{1}{2}q^3\!+\!\frac{5}{2}p^2q & 0 & pq^2 \\
\sm{2} & p^3\!+\!pq^2 & 0 & 3p^2q\!+\!q^3 & 0 & 2pq^2
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
A_{11} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & 0 & 0 & 0 & 0 & 0 \\
\sm{-1} & \frac{1}{2}p^2q\!+\!\frac{1}{2}q^3 & 0 & pq^2 & 0 & 0 \\
\sm{0} & 0 & p^2q\!+\!q^3 & 0 & 2pq^2 & 0 \\
\sm{1} & \frac{1}{2}pq^2\!+\!\frac{1}{2}p^3 & 0 & 2p^2q\!+\!q^3 & 0 & 2pq^2 \\
\sm{2} & p^3\!+\!pq^2 & 0 & 3p^2q\!+\!q^3 & 0 & 2pq^2
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
and
\begin{equation}
B = \left( \begin{array}{cc} B_{00} & B_{01} \\ B_{10} & B_{11} \end{array} \right)
\end{equation}
where
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
B_{00} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & 2p^2q & 0 & p^3\!+\!2pq^2 & 0 & 2p^2q \\
\sm{-1} & 0 & 0 & p^2q & 0 & pq^2 \\
\sm{0} & 0 & 2p^2q & 0 & 2pq^2 & 0 \\
\sm{1} & 2p^2q & 0 & p^3\!+\!2pq^2 & 0 & p^2q \\
\sm{2} & 0 & 0 & 0 & 0 & 0
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
B_{01} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & p^2q\!+\!q^3 & 0 & p^3\!+\!3pq^2 & 0 & 2p^2q \\
\sm{-1} & \frac{1}{2}p^3\!+\!\frac{1}{2}pq^2 & 0 & p^2q & 0 & 0 \\
\sm{0} & 0 & p^3\!+\!pq^2 & 0 & 2p^2q & 0 \\
\sm{1} & \frac{1}{2}q^3\!+\!\frac{1}{2}p^2q & 0 & p^3\!+\!2pq^2 & 0 & 2p^2q \\
\sm{2} & 0 & 0 & 0 & 0 & 0
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
B_{10} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & 0 & 0 & 0 & 0 & 0 \\
\sm{-1} & 0 & 0 & p^3 & 0 & p^2q \\
\sm{0} & 0 & 2p^3 & 0 & 2p^2q & 0 \\
\sm{1} & 2p^3 & 0 & 3p^2q & 0 & pq^2 \\
\sm{2} & 2p^3 & 0 & 4p^2q & 0 & 2pq^2
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
\vspace{1em}
\begin{IEEEeqnarray}{c}
\setlength\arraycolsep{0.2em}
\renewcommand{\arraystretch}{1.1}
B_{11} = \fharray{2em}{0em}{1em}{0em}{1em}{r@{\hspace{1.5em}}ccccc}{
& \sm{-2} &\sm{-1} & \sm{0} & \sm{1} & \sm{2} \\
\sm{-2} & 0 & 0 & 0 & 0 & 0 \\
\sm{-1} & \frac{1}{2}p^2q\!+\!\frac{1}{2}q^3 & 0 & pq^2 & 0 & 0 \\
\sm{0} & 0 & p^2q\!+\!q^3 & 0 & 2pq^2 & 0 \\
\sm{1} & \frac{1}{2}pq^2\!+\!\frac{1}{2}p^3 & 0 & 2p^2q\!+\!q^3 & 0 & 2pq^2 \\
\sm{2} & p^3\!+\!pq^2 & 0 & 3p^2q\!+\!q^3 & 0 & 2pq^2
} \IEEEeqnarraynumspace
\end{IEEEeqnarray}
\pagebreak
\section{Solving the recurrent equation}\label{sec:solving_the_recurrent_equation}
Consider the second equation in \eqref{eq: recurrent_equation}. It follows from \eqref{eq: viterbi_bit_error_probability_w} that we are only interested in the asymptotic values, and hence letting $t$ tend to infinity yields
\begin{IEEEeqnarray}{rCl}
\label{eq: b_infinity}
\bs{b}_{\infty} & = & \bs{b}_{\infty} \Pi
\end{IEEEeqnarray}
where $\bs{b}_{\infty}$ can be chosen as
\begin{IEEEeqnarray}{rCl}
\label{inf_case}
\bs{b}_{\infty} & = & (\bs{p}_{\infty}\, \bs{p}_{\infty} \ldots \bs{p}_{\infty})
\end{IEEEeqnarray}
To obtain the last equality, we took into account that $\Pi$ is a block-diagonal matrix whose diagonal elements are given by the transition probability matrix $\Phi$ which satisfies \eqref{steady-state}. Based on these observations, \eqref{eq: recurrent_equation} can be simplified to
\begin{IEEEeqnarray}{rCl}
\label{eq: simplified_recurrent_equation}
\bs{w}_{t + 1} & = & \bs{w}_{t} A + \bs{b}_{\infty} B
\end{IEEEeqnarray}
By iterating the recurrent equation \eqref{eq: simplified_recurrent_equation} and using the initial value \eqref{eq: initial_value_w}, the vector of the information weights at time instant $t+1$ is given by
\begin{IEEEeqnarray}{rCl}
\bs{w}_{t+1} & = & \bs{b}_{\infty} B A^t + \bs{b}_{\infty} B A^{t-1} + \cdots + \bs{b}_{\infty} B
\label{eq: w_t_plus_1}
\end{IEEEeqnarray}
Taking its limit, it follows that
\begin{IEEEeqnarray}{rCl}
\lim_{t \to \infty} \frac{\bs{w}_t}{tb} & = & \lim_{t \to \infty} \frac{\bs{w}_{t+1}}{tb} = \lim_{t \to \infty} \frac{1}{tb} \sum_{j=0}^t \bs{b}_{\infty} B A^{t-j} \nonumber \\
& = & \bs{b}_{\infty} B A^{\infty}/b \label{eq: final_limit}
\end{IEEEeqnarray}
where $A^{\infty}$ denotes the limit of the sequence $A^t$ when $t$ tends to infinity and we used the fact that, if a sequence converges to a finite limit, then it is Ces\`{a}ro-summable to the same limit.
From \eqref{eq:row_sum} it follows that
\begin{IEEEeqnarray}{rCl}
\label{eq: left_eigenvector}
\bs{e}_{\text{L}} & = & (\bs{p}_{\infty} \enspace \bs{p}_{\infty} \ldots \bs{p}_{\infty})
\end{IEEEeqnarray}
satisfies
\begin{IEEEeqnarray}{rCl}
\bs{e}_{\text{L}} A & = & \bs{e}_{\text{L}}
\end{IEEEeqnarray}
and hence $\bs{e}_{\text{L}}$ is a left eigenvector with eigenvalue $\lambda = 1$. Due to the nonnegativity of $A$, $\lambda = 1$ is a maximal eigenvalue of $A$ (Corollary 8.1.30 \cite{Horn1990}). Let $\bs{e}_{\text{R}}$ denote the right eigenvector corresponding to the eigenvalue $\lambda = 1$ normalized such that $\bs{e}_{\text{L}} \bs{e}_{\text{R}} = 1$. If we remove the allzero rows and corresponding columns from the matrix $A$ we obtain an irreducible matrix which has a unique maximal eigenvalue $\lambda = 1$ (Lemma 8.4.3~\cite{Horn1990}). Hence, it follows (Lemma 8.2.7, statement (i)~\cite{Horn1990}) that
\begin{IEEEeqnarray}{rCl}
\label{eq: multiplication_eigenvectors}
A^{\infty} = \bs{e}_{\text{R}} \bs{e}_{\text{L}}
\end{IEEEeqnarray}
Combining \eqref{eq: final_limit}, \eqref{eq: left_eigenvector}, and \eqref{eq: multiplication_eigenvectors} yields
\begin{IEEEeqnarray}{rCl}
\label{eq: combination}
\lim_{t \to \infty} \frac{\bs{w}_t}{tb} & = & \bs{b}_{\infty} B \bs{e}_{\text{R}} ( \bs{p}_{\infty} \enspace \bs{p}_{\infty} \ldots \bs{p}_{\infty}) / b
\end{IEEEeqnarray}
Following \eqref{eq: viterbi_bit_error_probability_w}, by summing up the first $M$ components of the vector $( \bs{p}_{\infty} \enspace \bs{p}_{\infty} \ldots \bs{p}_{\infty})$ on the right side of \eqref{eq: combination}, we obtain the closed form expression for the exact bit error probability as
\begin{IEEEeqnarray}{rCl}
\label{eq: final}
P_{\text{b}} & = & \bs{b}_{\infty} B \bs{e}_{\text{R}} / b
\end{IEEEeqnarray}
To summarize, the exact bit error probability $P_\text{b}$ for Viterbi decoding of a rate $R=b/c$ minimal convolutional encoder, when communicating over the BSC, is calculated as follows:
\begin{itemize}
\item Construct the set of metric states and find the stationary probability distribution $\bs{p}_{\infty}$.
\item Determine the matrices $A$ and $B$ as in \secref{recurrent_equation} and compute the right eigenvector $\bs{e}_{\text{R}}$ normalized according to $(\bs{p}_{\infty} \enspace \bs{p}_{\infty} \ldots \bs{p}_{\infty}) \bs{e}_{\text{R}} = 1$.
\item Calculate the exact bit error probability $P_{\text{b}}$ using \eqref{eq: final}.
\end{itemize}
For the encoder shown in \figref{ocf_realization} we obtain
\begin{IEEEeqnarray}{rCl}
P_{\text{b}} & = & \left(4p - 2p^2 + 67p^3 - 320p^4 + 818p^5 - 936p^6 - 884p^7 \right. \nonumber \\
& & \,\,+ 5592p^8 - 11232p^9 + 13680p^{10} - 11008p^{11} \nonumber \\
& & \left.\,\,+ 5760p^{12} - 1792p^{13} + 256p^{14} \right) / \left(2 - 5p + 41p^2 \right. \nonumber\\
& & \,\, - 128p^3 + 360p^4 - 892p^5 + 1600p^6 - 1904p^7 \nonumber \\
& & \left.\,\,+ 1440p^8 - 640p^9 + 128p^{10} \right) \nonumber \\
& = & 2p + 4p^2 + \frac{5}{2}p^3 - \frac{431}{4}p^4 - \frac{125}{8}p^5 + \frac{32541}{16}p^6 \nonumber \\
& & \, - \frac{70373}{32}p^7 - \frac{1675587}{64}p^8 + \frac{7590667}{128}p^9 \nonumber \\
& & + \frac{67672493}{256}p^{10} - \cdots
\end{IEEEeqnarray}
If we instead realize the \textit{minimal} generator matrix \eqref{eq:ocf_encoding_matrix_R23} in controller canonical form, we obtain a \textit{nonminimal} ($4$-state) encoder with $M=12$ normalized cumulative metric state; \textit{cf.},~the Remark after \eqref{eq: viterbi_bit_error_probability}. Its exact bit error probability is slightly worse than that of its minimal realization in observer canonical form.
\section{Some Examples}\label{sec:examples}
\begin{figure}[t]
\centering
\includegraphics{ebep_r12-external}
\vspace{-1em}
\caption{\label{fig:graph}Exact bit error probability for the rate $R=1/2$ minimal encoders of memory $m=1$ ($2$-state) $G(D)=(1 \quad 1+D)$, memory $m=2$ ($4$-state) $G(D)=(1+D^2 \quad 1+D+D^2)$, memory $m=3$ ($8$-state) $G(D)=(1+D^2+D^3 \quad 1+D+D^2+D^3)$, and memory $m=4$ ($16$-state) $G(D)=(1+D+D^4 \quad 1+D+D^2+D^3+D^4)$.}
\vspace{-1em}
\end{figure}
\begin{figure*}[t]
\centering
\includegraphics{trellis_m2-external}
\caption{\label{fig:four_trellis_sections_m2}Four different trellis sections of the in total $124$ for the $G(D) = (1+D^2 \quad 1+D+D^2)$ generator matrix.}
\vspace{-1em}
\end{figure*}
First we consider some rate $R=1/2$, memory $m=1,2,3$, and $4$ convolutional encoders; that is, encoders with $2, 4, 8$, and $16$ states, realized in controller canonical form. In \figref{graph} we plot the exact bit error probability for those four convolutional encoders.
\begin{example}
If we draw all $20$ trellis sections for the rate $R=1/2$, memory $m=1$ ($2$-state) convolutional encoder with generator matrix $G(D) = (1 \quad 1+D)$ realized in controller canonical form, we obtain the normalized cumulative metric states $\left\{ -2, -1, 0, 1, 2 \right\}$. Its metric state Markov chain yields the stationary probability distribution
{\setlength\arraycolsep{0.1em}
\begin{equation}
\bs{p}^\text{T}_{\infty} = \frac{1}{1+3p^2-2p^3}
\left(\begin{array}{ccccccccc}
1 & - & 4p & + & 8p^2 & - & 7p^3 & + & 2p^4 \\
& & 2p & - & 5p^2 & + & 5p^3 & - & 2p^4 \\
& & 2p & - & 3p^2 & + & 3p^3 \\
& & & & 2p^2 & - & 3p^3 & + & 2p^4 \\
& & & & p^2 & + & p^3 & - & 2p^4
\end{array}\right) \label{eq:stationary_memory1}
\end{equation}}%
Based on these $20$ trellis sections, the $10 \times 10$ matrices $A$ and $B$ are constructed as
\begin{equation}
A = \left(\begin{array}{cc} A_{00} & A_{01} \\ A_{10} & A_{11} \end{array}\right)
\end{equation}
and
\begin{equation}
B = \left(\begin{array}{cc} \bs{0}_{5,5} & A_{01} \\ \bs{0}_{5,5} & A_{11} \end{array}\right) \label{eq:ex_B_matrix}
\end{equation}
where $\bs{0}_{5,5}$ denotes the $5 \times 5$ all-zero matrix. The normalized right eigenvector of $A$ is
\begin{IEEEeqnarray}{rCl}
\label{eq:right_eigenvector_memory_1}
\bs{e}_{\text{R}} & = & \scriptscriptstyle
\left( \begin{array}{c}
0 \\ 0 \\ 0 \\ 0 \\ 0 \\ 0 \\ \displaystyle\frac{pq}{2} \\[0.7em] \displaystyle\frac{4pq}{2-p+4p^2-4p^3} \\[1em] \displaystyle\frac{(2 + 7p - 12p^2 + 13p^3 - 12p^4 + 4p^5)}{2(2-p+4p^2-4p^3)} \\[0.7em] 1
\end{array}\right)
\end{IEEEeqnarray}
Finally, inserting \eqref{eq:stationary_memory1}, \eqref{eq:ex_B_matrix}, and \eqref{eq:right_eigenvector_memory_1} into \eqref{eq: final} yields the following expression for the exact bit error probability
\begin{IEEEeqnarray}{rCl}
P_{\text{b}} & = & \frac{14p^2 - 23p^3 + 16p^4 + 2p^5 - 16p^6 + 8p^7}{(1+3p^2-2p^3)(2 - p + 4p^2 - 4p^3)} \nonumber \\
& = & 7p^2 - 8p^3 - 31p^4 + 64p^5 + 86p^6 - \frac{635}{2}p^7 \nonumber \\
& & \,- \frac{511}{4}p^8 + \frac{10165}{8}p^9 - \frac{4963}{16}p^{10} - \cdots
\end{IEEEeqnarray}
which coincides with the exact bit error probability formula given in \cite{Best1995}.
\end{example}
\begin{example}
For the rate $R=1/2$, memory $m=2$ ($4$-state) convolutional encoder with generator matrix $G(D) = (1+D^2 \quad 1+D+D^2)$ realized in controller canonical form, we obtain, for example, the four trellis sections for $\phi_t = \left( 000 \right)$ shown in \figref{four_trellis_sections_m2}. The corresponding metric states at times $t+1$ are $\phi_{t+1} = \left( -1\,0\,-\!\!1\right)$ and $\phi_{t+1} = \left( 1\,0\,1 \right)$.
Completing the set of trellis sections yields in total $M=31$ different normalized cumulative metric states, and hence the $124 \times 124$ matrices $A$ and $B$ have the following block structure
\pagebreak
\vphantom{.}\\\vspace{-0.5em}
\begin{equation}
A =
\left(\begin{array}{cccc}
A_{00} & \bs{0}_{31,31} & A_{02} & \bs{0}_{31,31} \\
A_{10} & \bs{0}_{31,31} & A_{12} & \bs{0}_{31,31} \\
\bs{0}_{31,31} & A_{21} & \bs{0}_{31,31} & A_{23} \\
\bs{0}_{31,31} & A_{31} & \bs{0}_{31,31} & A_{33}
\end{array}\right)
\end{equation}
\vspace{-0.5em}
and
\begin{equation}
B =
\left(\begin{array}{cccc}
\bs{0}_{31,31} & \bs{0}_{31,31} & A_{02} & \bs{0}_{31,31} \\
\bs{0}_{31,31} & \bs{0}_{31,31} & A_{12} & \bs{0}_{31,31} \\
\bs{0}_{31,31} & \bs{0}_{31,31} & \bs{0}_{31,31} & A_{23}\\
\bs{0}_{31,31} & \bs{0}_{31,31} & \bs{0}_{31,31} & A_{33}
\end{array}\right)
\end{equation}
Following the method for calculating the exact bit error probability described in \secref{solving_the_recurrent_equation} we obtain
\begin{IEEEeqnarray}{rCl}
P_{\text{b}} & = & 44p^3 + \frac{3519}{8}p^4 - \frac{14351}{32}p^5 - \frac{1267079}{64}p^6 \nonumber\\
& & \,- \frac{31646405}{512}p^7 + \frac{978265739}{2048}p^8 \nonumber\\
& & \, + \frac{3931764263}{1024}p^9 - \frac{48978857681}{32768}p^{10} - \cdots \label{eq:ebep_m2}
\end{IEEEeqnarray}
which coincides with the previously obtained result by Lentmaier et al.~\cite{Lentmaier2004}.
\end{example}
\begin{example}
For the rate $R=1/2$, memory $m=3$ ($8$-state) convolutional encoder with generator matrix $G(D) = (1+D^2+D^3 \quad 1+D+D^2+D^3)$ realized in controller canonical form we have $M=433$ normalized cumulative metric states and the $A$ and $B$ matrices are of size $433\cdot 2^3 \times 433 \cdot 2^3$.
Since the complexity of the symbolic derivations increases greatly, we can only obtain a numerical solution of \eqref{eq: final}, as shown in \figref{graph}.
\end{example}
\begin{example}
For the rate $R=1/2$, memory $m=4$ ($16$-state) convolutional encoder with generator matrix $G(D) = (1+D^2+D^3+D^4 \quad 1+D+D^4)$ realized in controller canonical form, we have as many as $M=188687$ normalized cumulative metric states. Thus, the matrices $A$ and $B$ are of size $188687 \cdot 2^4 \times 188687 \cdot 2^4$. The corresponding numerical solution of \eqref{eq: final} is plotted in \figref{graph}.
\end{example}
The obvious next step is to try a rate $R=1/2$, memory $m=5$ ($32$-state) convolutional encoder. We tried the generator matrix $G(D) = (1+D+D^2+D^3+D^4+D^5 \quad 1+D^3+D^5)$ realized in controller canonical form but were only able to show that the number of cumulative normalized metric states $M$ exceeds $4130000$.
\begin{figure}[!b]
\centering
\vspace{-1em}
\includegraphics{ebep_sys_m2-external}
\vspace{-1.5em}
\caption{\label{fig:equivalent_m2}Exact bit error probability for the rate $R=1/2$ memory $m=2$ minimal encoders with $G_1(D) = ( 1+D^2 \quad 1+D+D^2)$, $G_2(D) = (1 \quad (1+D^2)/(1+D+D^2))$, and $G_3(D) = (1 \quad (1+D+D^2)/(1+D^2))$.}
\end{figure}
\begin{example}
Consider the generator matrix $G_1(D) = ( 1+D^2 \quad 1+D+D^2 )$ and its equivalent systematic generator matrices $G_2(D) = (1 \quad (1+D^2)/(1+D+D^2))$ and $G_3(D) = (1 \quad (1+D+D^2)/(1+D^2))$. When realized in controller canonical form, all three realizations have $M=31$ normalized cumulative metric states. The exact bit error probability for $G_1(D)$ is given by \eqref{eq:ebep_m2}. For $G_2(D)$ and $G_3(D)$ we obtain
\begin{IEEEeqnarray}{rCl}
P_{\text{b}} & = & \frac{163}{2}p^3 + \frac{365}{2}p^4 - \frac{24045}{8}p^5 - \frac{1557571}{128}p^6 \nonumber\\
& & \,+ \frac{23008183}{512}p^7 + \frac{1191386637}{2048}p^8 \nonumber \\
& & \,+ \frac{4249634709}{8192}p^9 + \frac{132555764497}{8192}p^{10} - \cdots \label{eq:ebep_m2_5_7}
\end{IEEEeqnarray}
and\vspace{-0.5em}
\begin{IEEEeqnarray}{rCl}
P_{\text{b}} & = & \frac{141}{2}p^3 + \frac{1739}{8}p^4 - \frac{71899}{32}p^5 - \frac{1717003}{128}p^6 \nonumber \\
& & \,+ \frac{2635041}{128}p^7 + \frac{540374847}{1024}p^8 \nonumber \\
& & \,+ \frac{9896230051}{8192}p^9 - \frac{402578056909}{32768}p^{10} - \cdots \label{eq:ebep_m2_7_5}
\end{IEEEeqnarray}
respectively. The corresponding numerical results are illustrated in \figref{equivalent_m2}.
\end{example}
\begin{figure}[t]
\centering
\includegraphics{ebep_r23-external}
\vspace{-1em}
\caption{\label{fig:ebep_r23}Exact bit error probability for the rate $R=2/3$, overall constraint length $\nu=2,3$, and $4$ ($4$-state, $8$-state, and $16$-state, respectively) minimal encoders whose generator matrices are given in \tabref{encodingmatrices}.}
\end{figure}
\begin{example}
The exact bit error probabilities for the rate $R=2/3$ $4$-state, $8$-state, and $16$-state generator matrices, given in \tabref{encodingmatrices} and realized in controller canonical form, are plotted in \figref{ebep_r23}.
\begin{table}[!t]
\centering
\caption{\label{tab:encodingmatrices}Rate $R=2/3$ generator matrices}
\renewcommand{\arraystretch}{1.2}
\setlength\tabcolsep{0.2em}
\begin{tabular}{c|c|c|c}
$G(D)$ & \#states & $d_\text{free}$ & $M$ \\[0.5em] \shhline[1pt]
\multirow{3}{*}{\renewcommand{\arraystretch}{1}$\left(\begin{array}{ccc} D & 1+D & 1+D \\ 1 & D & 1+D \end{array}\right)$ } & \multirow{3}{*}{$4$} &\multirow{3}{*}{$3$} & \multirow{3}{*}{$19$} \\ &&\\ &&\\ \shhline
\multirow{3}{*}{\renewcommand{\arraystretch}{1}$\left(\begin{array}{ccc} 1+D & D & 1 \\ D^2 & 1 & 1+D+D^2 \end{array}\right)$ } & \multirow{3}{*}{$8$} &\multirow{3}{*}{$4$} & \multirow{3}{*}{$347$} \\ &&\\ &&\\ \shhline
\multirow{3}{*}{\renewcommand{\arraystretch}{1}$\left(\begin{array}{ccc} D+D^2 & 1 & 1+D^2 \\ 1 & D+D^2 & 1+D+D^2 \end{array}\right)$} & \multirow{3}{*}{$16$} &\multirow{3}{*}{$5$} & \multirow{3}{*}{$15867$} \\ &&\\ &&\\ \shhline[1pt]
\end{tabular}\vspace{0.5em}
\end{table}
As an example, the $4$-state encoder has the exact bit error probability
\begin{IEEEeqnarray}{rCl}
P_{\text{b}} & = & \frac{67}{2}p^2+ \frac{17761}{48}p^3 - \frac{2147069}{648}p^4 - \frac{1055513863}{46656}p^5 \nonumber\\
& & \,+ \frac{123829521991}{559872}p^6 + \frac{67343848419229}{60466176}p^7 \nonumber\\
& & \,- \frac{27081094434882419}{2176782336}p^8 - \frac{477727138796620247}{8707129344}p^9 \nonumber \\
& & \,+ \frac{1944829319763332473469}{2821109907456}p^{10} + \cdots
\end{IEEEeqnarray}
\end{example}
If we replace the BSC with the quantized additive white Gaussian noise (AWGN) channel, the calculation of the exact bit error probability follows the same method as described in \secref{solving_the_recurrent_equation}, but the computational complexity increases dramatically as illustrated by the following example.
\begin{figure}[t]
\centering
\includegraphics{ebep_quant-external}
\vspace{-1em}
\caption{\label{fig:ebep_quant}Exact bit error probability for the rate $R=1/2$, memory $m=2$ ($4$-state) encoder with $G(D)=(1+D^2 \quad 1+D+D^2)$ used to communicate over an AWGN channel with different quantization levels.}
\vspace{-0.8em}
\end{figure}
\begin{figure}[ht]
\centering
\includegraphics{quant_levels-external}
\vspace*{1mm}
\caption{\label{fig:quant_levels}Examples of uniform and Massey quantizations for an AWGN channel with SNR = $0 \text{dB}$.}
\vspace{-1em}
\end{figure}
\begin{example}
Consider the generator matrix $G(D) = (1+D^2 \quad 1+D+D^2)$ used to communicate over a quantized AWGN channel. We use different quantization methods, namely, uniform quantization \cite{Heller1971, Onyszchuk1993} and Massey quantization \cite{Massey1974, Johannesson1999}; see \figref{quant_levels}.
The uniform intervals were determined by optimizing the cut-off rate $R_0$. The Massey quantization thresholds $T_i$ between intervals were also determined by optimizing $R_0$, but allowing for nonuniform intervals. The realization in controller canonical form yields that, for all signal to noise ratios (SNRs), $E_\text{b} / N_0$, and uniform quantization with $7$, $8$, and $9$ levels, the number of the normalized cumulative metric states is $M=1013$, $M=2143$, and $M=2281$, respectively. However, for the Massey quantization the number of normalized cumulative metric states varies with both the number of levels and the SNR. Moreover, these numbers are much higher. For example, considering the interval between $0\,\text{dB}$ and $3.5\,\text{dB}$ with $8$ quantization levels, we have $M=16639$ for $E_\text{b} / N_0 \leq 2.43\,\text{dB}$, while for $E_\text{b} / N_0 > 2.43\,\text{dB}$ we obtain $M=17019$. The exact bit error probability for this $4$-state encoder is plotted for all different quantizations in \figref{ebep_quant}, ordered from worst (top) to best (bottom) as
{\vspace{0.5em}\setlength\tabcolsep{0em}
\begin{tabular}{rll}
\quad (i) &\, Uniform~~& $8$ levels \\
\quad (ii) &\, Uniform~~& $7$ levels \\
\quad (iii) &\, Massey ~~& $7$ levels \\
\quad (iv) &\, Uniform~~& $9$ levels \\
\quad (v) &\, Massey ~~& $8$ levels \\
\quad (vi) &\, Massey ~~& $9$ levels \\
\end{tabular}\vspace{0.5em}}%
All differences are very small, and hence it is hard to distinguish all the curves. It is interesting to notice that using $7$ instead of $8$ uniform quantization levels yields a better bit error probability. However, this is not surprising since the presence of a quantization bin around zero typically improves the quantization performance. Moreover, the number of cumulative normalized metric states for $7$ quantization levels is only about one half of that for $8$ quantization levels. Notice that such a subtle comparison of channel output quantizers has only become possible due to the closed form expression for the exact bit error probability.
\end{example}
\section{Conclusion}
We have derived a closed form expression for the exact bit error probability for Viterbi decoding of convolutional codes using a recurrent matrix equation. In particular, the described method is feasible to evaluate the performance of encoders with as many as $16$ states when communicating over the BSC. By applying our new approach to a $4$-state encoder used to communicate over the quantized AWGN channel, the expression for the exact error probability for Viterbi decoding is also derived. In particular, it is shown that the proposed technique can be used to select the optimal encoder implementation as well as the optimal channel output quantizer based on comparing their corresponding exact bit decoding error probability.
|
Subsets and Splits